You are on page 1of 676

Douglas-fir

The Genus Pseudotsuga

Denis P. Lavender and Richard K. Hermann


Douglas-fir: The Genus Pseudotsuga
Douglas-fir
The Genus Pseudotsuga
Denis P. Lavender and Richard K. Hermann

Forest Research Laborator y


The opinions expressed are those of the authors are do not necessarily represent those of the
Oregon Forest Research Laboratory or Oregon State University.

Copyright © 2014 by Denis P. Lavender and Richard K. Hermann


ISBN: 978-0-615-97995-3
Printed in the United States of America
Disclaimer
Mention of trade names and products does not constitute endorsement, recommendation for use, or promo-
tion of the products by the authors or the organizations with which they are affiliated.
The Oregon Forest Research Laboratory
The OSU College of Forestry Oregon Forest Research Laboratory (FRL), established by the Oregon
Legislature, is Oregon’s principal forest-related research engine. Faculty and students have been
providing science-based
knowledge, developing new technologies, and creating innovative decision support tools for more than
100 years. The breadth and depth of our faculty expertise enables us to strengthen fundamental
under- standing of forested ecosystems, help forest-based businesses compete in a global
marketplace, support the viability of communities, and inform public policy that balances
environmental protection and economic development. Research results are made available through
scientific journals, educational programs, and through FRL publications such as this, which are
directed as appropriate to forest landowners and managers, manufacturers and users of forest
products, leaders of government and industry, the scientific community, the conservation
community, and the general public.
Open Access
In keeping with a 100+ year history of making research publications available on request, the OSU College
of Forestry supports an Open Access policy. This book and other College of Forestry publications are
available through ScholarsArchive@OSU, a project of the OSU Libraries. The fully searchable pdf version
of Douglas-fir: The Genus Pseudotsuga, by Denis P. Lavender and Richard K. Hermann, is available online
and in color here: http://hdl.handle.net/1957/47168.

Cover photo: Douglas-fir (Pseudotsuga menziesii) cones. © Bryan Bernart Photography. 2014. All Rights Reserved.
Produced by
The Forestry Communications Group
Oregon State University
202 Peavy Hall
Corvallis, OR 97331
Voice: (541) 737-4270
Email: ForestryCommunications@oregonstate.edu
Editing, design, and layout by Caryn M. Davis, College of Forestry, Oregon State University.
Abstract
Lavender, Denis P., and Richard K. Hermann. 2014. Douglas-fir: The Genus Pseudotsuga.
Oregon Forest Research Laboratory, Oregon State University, Corvallis.

Douglas-fir (Pseudotsuga menziesii) has a wide distribution in North America and is one of
the tree species most widely distributed outside its natural range. The species has been
introduced to Europe, New Zealand, South America, and elsewhere around the world. At
present, Douglas-fir is an accepted and integral part of forest management in many
countries because of its economic importance and its reputation as a species that may be
able to deal with climate change.

This book provides an overview of research activities and findings that highlight unique aspects
of Douglas-fir physiology, genetics, and other related issues. It begins with the evolutionary
history and distribution of Douglas-fir and provides a detailed description of introductions of
Douglas-fir to other countries, including information about initial plantings, provenance trials,
and genetic tree improvement activities.

The sections about life history, drawn from extensive research and teaching experiences, include
detailed descriptions of flowering, seeds, root, and seedling physiology, followed by sections
about mycorrhizae and insects, diseases, and other biotic factors. It discusses research that
demonstrates some of the unique aspects of Douglas-fir physiology, for example: (1) Douglas-
fir has an annual growth cycle that includes a cold period in the late fall or early winter. Failure
to experience these low temperatures results in a substantial loss of vigor; (2) the reproductive
system of the stem is stimulated by material from the roots; and (3) the root system plays a
supportive role. Nutrient and moisture uptake are mediated by mycorrhizae.

This book is intended as a resource for everyone interested in understanding the opportunities
and challenges of managing Douglas-fir in a variety of regions and settings. It provides
information for historians and social scientists investigating forestry trends; researchers,
educators, and managers looking for detailed information in areas such as genetics and
regeneration practices; and all others interested in the beautiful trees we call Douglas-fir.
Contents
List of Figures...................................................................................................................................................xv
List of Tables....................................................................................................................................................xvi
Foreword................................................................................................................................................................xvii
Preface.....................................................................................................................................................................xix
About the Authors.............................................................................................................................................xxi
Acknowledgements..........................................................................................................................................xxii
1. Evolutionary History
Richard K. Hermann.........................................................................................................................................1
The Genus Pseudotsuga......................................................................................................................................1
Differentiation into varieties........................................................................................................3
Intraspecific variation..................................................................................................................4
Terpenes..................................................................................................................................6
Enzymes............................................................................................................................................7
Random amplified polymorphic DNA..................................................................................12
2. Natural Range
Richard K. Hermann.......................................................................................................................................15
History............................................................................................................................................ 15
Varieties of Pseudotsuga menziesii..................................................................................................................15
Range....................................................................................................................................................17
Altitudinal distribution...............................................................................................................21
Area occupied by Douglas-fir....................................................................................................22
Climate..................................................................................................................................................23
Soils.............................................................................................................................................................24
Associated forest cover..............................................................................................................25
Pseudotsuga macrocarpa..................................................................................................................................26
Pseudotsuga japonica........................................................................................................................................27
The Chinese Douglas-firs................................................................................................................28
Pseudotsuga wilsoniana..............................................................................................................................28
Pseudotsuga sinensis...................................................................................................................................29
Pseudotsuga gaussenii.................................................................................................................................29
Pseudotsuga forrestii...................................................................................................................................30
Pseudotsuga brevifolia................................................................................................................................30
3. Areas of Introduction
Richard K. Hermann.......................................................................................................................................31
Northern Hemisphere......................................................................................................................31
Western North America..................................................................................................................31
Alaska...................................................................................................................................................31
Hawaii........................................................................................................................................31
Central and Eastern North America.................................................................................................32
Europe.............................................................................................................................................33
Great Britain.........................................................................................................................................33
Ireland........................................................................................................................................ 39
Western Central Europe..................................................................................................................40
Germany.................................................................................................................................... 40
The Netherlands.........................................................................................................................45
Belgium.................................................................................................................................................46
Luxembourg...............................................................................................................................46
France....................................................................................................................................................47
Switzerland...........................................................................................................................................49
Austria........................................................................................................................................ 50
vii

Northern Europe..............................................................................................................................50
Denmark.................................................................................................................................... 50
Norway...................................................................................................................................... 51
Sweden.......................................................................................................................................52
Finland....................................................................................................................................... 53
Mediterranean Europe.....................................................................................................................54
Portugal......................................................................................................................................54
Spain.....................................................................................................................................................54
Italy.......................................................................................................................................................55
Croatia........................................................................................................................................56
Slovenia.................................................................................................................................................56
Greece...................................................................................................................................................56
Turkey........................................................................................................................................ 56
Cyprus........................................................................................................................................56
Eastern Europe................................................................................................................................ 57
Czech Republic.....................................................................................................................................57
Slovakia.................................................................................................................................................58
Hungary..................................................................................................................................... 59
Serbia....................................................................................................................................................60
Bosnia and Herzegovina............................................................................................................60
Bulgaria.................................................................................................................................................60
Romania................................................................................................................................................61
Poland........................................................................................................................................ 62
Baltic States................................................................................................................................................63
Lithuania.................................................................................................................................... 63
Latvia....................................................................................................................................................63
Estonia...................................................................................................................................................64
Former Soviet Union.......................................................................................................................64
Belarus..................................................................................................................................................64
Ukraine...................................................................................................................................... 64
Russia....................................................................................................................................................65
Asia.............................................................................................................................................................66
India........................................................................................................................................... 66
Sri Lanka...............................................................................................................................................66
Southern Hemisphere......................................................................................................................66
Southwestern Pacific............................................................................................................................66
New Zealand.............................................................................................................................. 66
Australia.....................................................................................................................................69
South America.................................................................................................................................71
Chile......................................................................................................................................................71
Argentina................................................................................................................................... 72
Africa..........................................................................................................................................................72
Republic of South Africa...........................................................................................................72
4. Provenance Trials
Richard K. Hermann.......................................................................................................................................73
Pacific Northwest Studies...............................................................................................................73
The 1954 Oregon State University provenance study................................................................74
The British Columbia Forest Service study................................................................................75
The University of British Columbia trial....................................................................................78
Short-term seedling tests............................................................................................................78
Common garden studies: variety glauca.....................................................................................81
The IUFRO International Douglas-fir Provenance Study................................................................83
Pre-IUFRO European Studies....................................................................................................83
German provenance trials...........................................................................................................83
viii

The 1904–1911 Fürstenberg trial..........................................................................................83


The 1910 Schwappach trial...................................................................................................84
The 1912 Kaiserslautern trial................................................................................................84
The 1930s trials...............................................................................................................................85
The 1932–1933 Wiedemann trials........................................................................................85
The 1930 Geyr von Schweppenburg trial..............................................................................87
The 1930 Fabricius trial..................................................................................................................87
The 1954/1958 Lower Saxony Forest Experiment Station trials...........................................87
The 1955/1958 Baden-Wuerttemberg trials..........................................................................88
The 1958 Schober trial....................................................................................................................88
The 1958 Hessian Forest Experiment Station trial................................................................89
The 1962–1963 Schmalenbeck Institute of Forest Genetics & Forest Tree Breeding trial....89
The 1961 German Democratic Republic trial........................................................................89
The IUFRO Trials.................................................................................................................................90
The 1970 Hesse trial.............................................................................................................90
The GDR IUFRO Trial...................................................................................................................91
Seed sources for Germany....................................................................................................91
International Studies........................................................................................................................92
British provenance trials............................................................................................................92
Irish provenance trials................................................................................................................94
Dutch provenance trials..............................................................................................................95
Belgian provenance trials...........................................................................................................96
French provenance trials............................................................................................................96
The AFOCEL provenance tests.............................................................................................98
Recommendations based on provenance tests.......................................................................99
Spanish provenance trials...........................................................................................................99
Italian provenance trials...........................................................................................................100
Austrian provenance trials........................................................................................................102
Bulgarian provenance trials......................................................................................................102
Bosnia and Herzegovina provenance trials...............................................................................103
Czech provenance trials...........................................................................................................104
Slovac provenance trials...........................................................................................................105
Hungarian provenance trials.....................................................................................................105
Polish provenance trials...........................................................................................................105
Danish provenance trials..........................................................................................................106
Norwegian provenance trials....................................................................................................108
Swedish provenance trials........................................................................................................108
Finnish provenance trials.........................................................................................................110
Latvian provenance trials.........................................................................................................110
Estonian provenance trials.......................................................................................................110
Turkish provenance trials.........................................................................................................111
Taiwan provenance trials..........................................................................................................111
New Zealand provenance trials................................................................................................112
Fort Bragg strain...........................................................................................................................114
Ashley strain.......................................................................................................................114
Beaumont strain..................................................................................................................114
Kaingaroa strain..................................................................................................................114
Australian provenance trials.....................................................................................................114
What has been learned from provenance experiments?.................................................................115
Survival...............................................................................................................................................115
Height growth..........................................................................................................................116
Southern hemisphere................................................................................................................116
Progeny of introduced populations...........................................................................................116
Conclusion................................................................................................................................................116
ix

5. Tree Breeding and Improvement


Richard K. Hermann......................................................................................................................................117
Genecology of Douglas-fir......................................................................................................................118
Quantitative genetics and inheritance.......................................................................................118
Heritabilities and amounts of genetic variation.........................................................................119
Genetic correlations..................................................................................................................119
Estimates of genetic gain..........................................................................................................120
Breeding Goals and Objectives......................................................................................................121
Primary breeding objectives.....................................................................................................121
Secondary breeding objectives.................................................................................................121
Steps in Tree Breeding Programs...................................................................................................122
Breeding zones......................................................................................................................... 122
Selection..............................................................................................................................................122
Genetic Improvement Programs....................................................................................................122
North America..........................................................................................................................122
Northwest Tree Improvement Program (NWTIC)...............................................................122
Weyerhaeuser Company......................................................................................................123
British Columbia Ministry of Forests (BCMP)....................................................................123
Inland Empire Tree Improvement Program (IETIC)...........................................................123
Europe...................................................................................................................................... 123
New Zealand.............................................................................................................................124
6. Flowering
Denis P. Lavender...........................................................................................................................................125
History and Nomenclature.............................................................................................................125
The role of juvenility and maturity in flowering.......................................................................127
Incidence of flowering..............................................................................................................130
Methodology of flower induction.............................................................................................131
Moisture stress.....................................................................................................................131
Light intensity......................................................................................................................131
Photoperiod.........................................................................................................................132
Fertilizers (mineral nutrition)..............................................................................................132
Temperature..............................................................................................................................133
Shoots............................................................................................................................................133
Root temperature.................................................................................................................134
Cultural treatments...................................................................................................................137
Girdling...............................................................................................................................137
Top pruning and branch thinning...................................................................................139
Root pruning and grafting...................................................................................................139
Gravimorphism, shading.....................................................................................................140
Plant growth regulators.............................................................................................................140
Gibberellins...................................................................................................................................140
Sex expression...............................................................................................................................143
Auxins................................................................................................................................. 143
Cytokinins......................................................................................................................................143
Abscisic acid..................................................................................................................................143
Carbohydrates......................................................................................................................144
Arginine...............................................................................................................................144
Polyamines.....................................................................................................................................144
Summary.................................................................................................................................. 144
Embryogeny..............................................................................................................................................145
Microsporangiate strobilus.......................................................................................................148
The pollen cone........................................................................................................................150
Ovulate strobilus.......................................................................................................................155
Pollen physiology..................................................................................................................... 157
x

Storage...........................................................................................................................................157
Viability of pollen.........................................................................................................................157
Fertilization.........................................................................................................................................159
Pollen distribution....................................................................................................................162
Pollination...........................................................................................................................................163
7. Seeds
Denis P. Lavender..........................................................................................................................................165
Seed Dormancy.............................................................................................................................165
Biochemistry and Seed Dormancy Breaking.................................................................................167
Growth regulators....................................................................................................................169
Gibberellins...................................................................................................................................169
Abscisic acid (ABA)......................................................................................................................170
Germination.................................................................................................................................. 170
Light....................................................................................................................................................172
Moisture...................................................................................................................................173
Temperature.............................................................................................................................175
Seed Tests.................................................................................................................................................177
Germination tests.....................................................................................................................177
Standard germination test....................................................................................................179
Cutting test..........................................................................................................................180
Biochemical quick tests.................................................................................................................180
Hydrogen peroxide..........................................................................................................180
Excised embryo.........................................................................................................................181
Seed-vigor tests........................................................................................................................181
Stratification........................................................................................................................................182
Cone and Seed Production............................................................................................................185
Seed size and anatomy.............................................................................................................185
Seed flight...........................................................................................................................................188
Seed size and germination........................................................................................................189
Seed development....................................................................................................................190
Seed processing (damage)........................................................................................................193
Seed storage.............................................................................................................................193
Storage under controlled conditions....................................................................................194
Storage under natural conditions.........................................................................................195
8. Seedlings
Denis P. Lavender..........................................................................................................................................197
Seedling Dormancy.......................................................................................................................197
Environment and dormancy initiation......................................................................................199
Dormancy and physiological response.....................................................................................199
Plant growth regulators and dormancy...............................................................................199
Dormancy Breaking......................................................................................................................200
Climate Change and Chilling........................................................................................................201
Dormancy and Growth Potential...................................................................................................202
Stress resistance and cold storage.............................................................................................202
Cold storage with light........................................................................................................203
Cold storage without light...................................................................................................203
Chlorophyll fluorescence.........................................................................................................204
Dormancy and the Concentration of Inorganic and Organic Constituents.....................................204
9. Cone and Seed Insects and Diseases
Denis P. Lavender..........................................................................................................................................207
Cone and Seed Insects...................................................................................................................207
Barbara colfaxiana....................................................................................................................................208
Leptoglossus occidentalis..........................................................................................................................209
xi
Megastigmus spermotrophus.....................................................................................................................210
Contarinia oregonensis.............................................................................................................................213
Contarinia washingtonensis......................................................................................................................214
Dioryctria abietella...................................................................................................................................215
Lepesoma lecontei.....................................................................................................................................216
Choristoneura occidentalis........................................................................................................................216
Diapause........................................................................................................................................216
Control.......................................................................................................................................... 218
Surveys................................................................................................................................................218
Artificial control.................................................................................................................................218
Systemic and chemical........................................................................................................218
Chemical attractants............................................................................................................219
Phenology................................................................................................................................ 219
Cold water spray.................................................................................................................219
Barriers..........................................................................................................................................220
Seed treatment.....................................................................................................................220
Natural control.........................................................................................................................220
Competition........................................................................................................................220
Biological control..........................................................................................................................221
Diseases....................................................................................................................................................222
Fungi........................................................................................................................................222
10. Roots
Denis P. Lavender................................................................................................................................. 225
Characteristics..........................................................................................................................................225
Function....................................................................................................................................................226
Relevance of Root Systems for Seedling Survival and Growth.....................................................227
Root regeneration potential......................................................................................................227
Root electrolyte leakage...........................................................................................................229
Temperature (heat)...................................................................................................................230
Ambient soil temperature....................................................................................................232
Soil moisture.......................................................................................................................................232
Light....................................................................................................................................................233
Organic residues................................................................................................................. 233
Root growth periodicity......................................................................................................234
11. Mycorrhizae
Denis P. Lavender................................................................................................................................. 235
Ectomycorrhizae.......................................................................................................................................235
Anatomy...................................................................................................................................236
Mycorrhizae..............................................................................................................................................237
Summary.............................................................................................................................239
Ecology...............................................................................................................................................239
Disease................................................................................................................................................240
Mycorrhizosphere......................................................................................................................... 241
Nutrient and water uptake...................................................................................................241
Physiology..........................................................................................................................................242
Soil.............................................................................................................................................................242
Fire......................................................................................................................................................243
Cost-Benefits of Mycorrhizae..................................................................................................................243
Conclusion................................................................................................................................................243
12. Adverse Abiotic Factors
Richard K. Hermann............................................................................................................................. 245
Frost..........................................................................................................................................................245

xii
Identification of frost injury.....................................................................................................245
Needles............................................................................................................................... 245
Buds...............................................................................................................................................246
Trunk.................................................................................................................................. 246
Roots..............................................................................................................................................246
Consequences of frost injury....................................................................................................246
Needles............................................................................................................................... 246
Buds...............................................................................................................................................246
Reproductive buds..............................................................................................................248
Stem tissue..........................................................................................................................248
Roots..............................................................................................................................................249
Cold acclimation and deacclimation.........................................................................................249
Assessment of cold hardiness...................................................................................................250
Age......................................................................................................................................................251
Genetics of cold hardiness..................................................................................................................251
Fall cold hardiness................................................................................................................... 251
Quantitative genetics................................................................................................................252
Coastal Douglas-fir.............................................................................................................252
Interior Douglas-fir.............................................................................................................254
Frost hardiness prediction models............................................................................................257
Interior Douglas-fir.............................................................................................................257
Coastal Douglas-fir.............................................................................................................257
Geographic Variation...............................................................................................................258
Aspect of site............................................................................................................................260
Frost-induced drought..............................................................................................................260
Nutrients...................................................................................................................................261
Intraspecific hybridization........................................................................................................262
Interspecific hybridization..................................................................................................................264
Frost heaving............................................................................................................................264
Snow, Ice, and Hail.......................................................................................................................264
Drought......................................................................................................................................... 265
Fire............................................................................................................................................................268
Deterioration of fire-killed Douglas-fir....................................................................................270
Air Pollutants................................................................................................................................ 271
Sulfur dioxide...........................................................................................................................271
Fluorides.................................................................................................................................. 271
Wind..............................................................................................................................................272
13. Ontogeny
Denis P. Lavender................................................................................................................................. 273
Ontogeny – Growth.......................................................................................................................273
Foliage, anatomy, quantity, distribution...................................................................................273
Seedlings.......................................................................................................................................273
Mature trees........................................................................................................................273
Branches..............................................................................................................................................273
Saplings.........................................................................................................................................273
Old growth..........................................................................................................................273
Ontogeny – Height........................................................................................................................ 274
Seedlings.............................................................................................................................................274
Saplings...............................................................................................................................................274
Mature trees............................................................................................................................. 274
Old growth...............................................................................................................................274
Ontogeny – Phenology..................................................................................................................275
Seedlings.............................................................................................................................................275
Saplings...............................................................................................................................................276

xiii
Ontogeny – Photosynthesis...........................................................................................................276
Young trees..............................................................................................................................276
Mature trees............................................................................................................................. 277
Competition..............................................................................................................................277
Ontogeny – Insects........................................................................................................................278
Seedlings.............................................................................................................................................278
Conifer seedling weevil......................................................................................................278
Black army cutworm...........................................................................................................278
Mature trees............................................................................................................................. 278
Western spruce budworm....................................................................................................278
Douglas-fir tussock moth....................................................................................................279
Western hemlock looper.....................................................................................................280
Douglas-fir beetle..........................................................................................................................280
Ontogeny – Disease.......................................................................................................................280
Seedlings.............................................................................................................................................280
Young trees..............................................................................................................................281
Phellinus weirii....................................................................................................................................281
Saplings, young growth............................................................................................................281
Armillaria root...............................................................................................................................281
Mature trees, old growth..........................................................................................................281
Root and butt rots................................................................................................................281
Young stands............................................................................................................................281
Phellinus weirii....................................................................................................................................282
Swiss needle cast...........................................................................................................................282
Mature stands, old growth..................................................................................................282
Swiss needle cast...........................................................................................................................283
Fomes pini............................................................................................................................................283
References..............................................................................................................................................................285
Index.......................................................................................................................................................................347

xiv
Figures
Figure 1.1 Frothingham’s (1990) division of the range of Douglas-fir into five growing regions...............................4
Figure 1.2 Chemical races of Douglas-fir based on leaf oil terpenes...........................................................................6
Figure 1.3 Geographic differentiation of Douglas-fir based on cortical monoterpenes...............................................8
Figure 1.4 Locations of seed from 104 sources used by Li and Adams (1989) in their allozyme study........................9
Figure 1.5 Genetic distances between provenances used for the Li and Adams (1989) allozyme study.....................11
Figure 1.6 Seed sources matched with populations from Li and Adams (1989) allozyme study.................................12
Figure 1.7 Ecotypes of Douglas-fir based on geographic variation of allelic structures (from Klumpp 1999)..........13
Figure 2.1 Lambert’s (1832) botanical illustration of Douglas-fir.............................................................................15
Figure 2.2 Natural range of Douglas-fir......................................................................................................................17
Figure 2.3 Line designates the northern limit of Douglas-fir in British Columbia.....................................................18
Figure 2.4 Range map of Douglas-fir in British Columbia..........................................................................................20
Figure 2.5 Range of Douglas-fir in Mexico (modified from Martinez 1963)...............................................................21
Figure 2.6 Range of Pseudotsuga macrocarpa (from Minnich 1982).........................................................................27
Figure 2.7 Range of Pseudotsuga japonica Beissn. (from Hayashi 1952)..................................................................28
Figure 2.8 Range of Pseudotsuga in China (from Chengde 1981)..............................................................................29
Figure 3.1 Areas ocupied by Douglas-fir in Europe, in hectares................................................................................34
Figure 3.2 Douglas-fir stands in Luxembourg (from Decker 1965)............................................................................47
Figure 3.3 Standing volume (m3) of Douglas-fir in France........................................................................................48
Figure 3.4 Locations of Douglas-fir stands in Norway................................................................................................51
Figure 3.5 Douglas-fir stands in Sweden, totaling 100 ha...........................................................................................52
Figure 3.6 Douglas-fir stands in Finland.....................................................................................................................53
Figure 3.7 Area stocked with Douglas-fir in Czech Republic......................................................................................58
Figure 3.8 Location of oldest Douglas-fir stands in Hungary.....................................................................................59
Figure 3.9 Douglas-fir stands in Romania...................................................................................................................61
Figure 3.10 Douglas-fir stands in Poland....................................................................................................................63
Figure 3.11 Douglas-fir in Lithuania...........................................................................................................................64
Figure 3.12 Potential volume growth of Douglas-fir in areas of Chile.......................................................................71
Figure 4.1 Seed collection and planting sites of 1912 Douglas-fir heredity study......................................................74
Figure 4.2 Collection sites for 1966 British Columbia Forest Service provenance study...........................................76
Figure 4.3 Sites of a series of four provenance plantations established 1968 to 1975................................................77
Figure 4.4 IUFRO seed lots used in British provenance trials grouped into 10 zones...............................................93
Figure 4.5 Locations of Douglas-fir provenance and progeny tests in France...........................................................97
Figure 4.6 AFOCEL provenance test plantations established by AFOCEL in France...............................................98
Figure 4.7 Location of Spanish Douglas-fir provenance tests (from Hernandez et al. 1993).....................................99
Figure 4.8 Seed zones with provenances recommended for Austria..........................................................................103
Figure 4.9 Location of Bulgarian provenance test plantations (squares) and cities.................................................103
Figure 4.10 Collection sites for Swedish provenance tests........................................................................................109
Figure 4.11 Sites of Egon Larsen seed collection recommended for use in New Zealand........................................113
Figure 4.12 Location of six plantations of the 1959 provenance test in New Zealand..............................................113
Figure 4.13 Australian Douglas-fir provenance plantations......................................................................................114
Figure 6.1 Seed and pollen development...................................................................................................................126
Figure 10.1 Root growth periodicity in Douglas-fir..................................................................................................234
Figure 12.1 Course of relative frost hardiness in four single tree progenies............................................................252
Figure 12.2 Frost hardiness and injury (from Rehfeldt)............................................................................................257
Figure 12.3 Percentage of total log volume of mature Douglas-fir decayed after fire-kill.......................................270

xv
Tables
Table 1.1 Racial differentiation, genetic distances, and genic and allelic diversity....................................................13
Table 2.1 Individual tree characteristics for Douglas-fir in the western United States...............................................16
Table 2.2 Area of Douglas-fir in the United States by variety and date of inventory...................................................23
Table 2.3 Area occupied by Douglas-fir 1987 in the United States by region and productivity class.........................23
Table 2.4 Volume of standing Douglas-fir timber in the United States.......................................................................23
Table 2.5 Climatic data for five subdivisions of the range of Douglas-fir...................................................................24
Table 3.1 Area stocked with Douglas-fir, according to census of woodlands in 1947 and 1982.................................37
Table 3.2 Area and standing volume of Douglas-fir in Great Britain by planting year classes...................................37
Table 3.3 Standing volume of Douglas-fir in Great Britain by size classes (from Locke 1987)...................................38
Table 3.4 Area occupied by Douglas-fir and mean yield class in state plantations in Ireland by age.........................39
Table 3.5 Growth and productivity data of Douglas-fir in the Ukraine (from Brodovich 1967).................................65
Table 3.6 Number of hectares stocked with Douglas-fir in New Zealand in 1974.......................................................66
Table 3.7 Number of hectares stocked with Douglas-fir in New Zealand, 1940 to 1991.............................................67
Table 3.8 Douglas-fir age class distribution in New Zealand as of April 1, 1991)......................................................67
Table 3.9 Imports of Douglas-fir seed to New Zealand since 1926 (from Wilcox 1978)..............................................68
Table 3.10 Sources and quantities of Douglas-fir seed used in New Zealand 1926-1974...........................................69
Table 4.1 Provenances and test locations in the 1932-1933 Wiedemann trials...........................................................85
Table 4.2 Provenances and test locations in the 1932-1933.........................................................................................86
Table 4.3 Letter codes used by to indicate elevation of seed source and number of frost-free days............................87
Table 4.4 Performance of seed in test plantations in 23 Bavarian forest districts.......................................................87
Table 4.5 Provenances in five Danish trials from 1930 to 1968.................................................................................107
Table 4.6 Locations of the origin of four provenances used in Taiwan trials.............................................................111
Table 4.7 Superior Douglas-fir provenances selected for use in New Zealand (from Wilcox 1974).........................112
Table 6.1 Effect of seedling root temperatures on developing reproductive structures in Douglas-fir.....................135
Table 6.2 Effect of seedling root temperatures on the growth of vegetative shoots in Douglas-fir............................135
Table 6.3 Effect of seedling root temperature upon the mineral content of scion and seedling tissue.......................135
Table 6.4 Effect of seedling root temperature upon plant moisture stress in Douglas-fir..........................................135
Table 7.1 Change of weight, nitrogen, and MC of developing Douglas-fir cone scale and seed...............................172
Table 7.2 Average percentage of germination without prechilling and with various periods....................................178
Table 7.3 Average germination by years using no chilling or prechilling of Douglas-fir seed..................................178
Table 7.4 Comparison of germination results by range and treatment......................................................................178
Table 7.5 General rating of Douglas-fir cone crops in Washington and Oregon from 1909 to 1954.......................187
Table 7.6 The following sequence of seed years noted in The Netherlands 1931–67................................................187
Table 7.7 External changes in Douglas-fir cones during the period of seed maturation...........................................193
Table 8.1 Historical nomenclature of dormancy phenomena.....................................................................................198
Table 8.2 Ecodormany, paradormancy, and endodormancy......................................................................................199
Table 8.3 Effects of a daily photoperiod during cold storage upon the growth responses of seedlings.....................203
Table 8.4 Storage condition and survival for seedlings lifted in October..................................................................203
Table 8.5 Concentration of constituents in root pressure exudates from seedlings...................................................205
Table 8.6 Sugars and amino acids in root pressure exudates of conifer seedlings....................................................205
Table 10.1 Root growth comparison between rye and Douglas-fir............................................................................226
Table 10.2 Relation between ovendry weights to thermoperiods and soil temperatures............................................231
Table 12.1 Percentage of trees damaged by the February 1956 freeze in eastern Germany.....................................251
Table 12.2 Percent survival of progeny from mother trees in the first 8 years after outplanting..............................252
Table 12.3 Performance of Douglas-fir provenances within 8 groups at the Kellogg plantation..............................259
Table 12.4 Volume lost in three Oregon Douglas-fir stands to ice damage in January 1942....................................265

xvi
Foreword
Experience is the mother of knowledge. —English proverb

Denis Lavender and Richard Hermann combine more than 60 years of forestry
research experience each in completing this synthesis about Douglas-fir. This
monumental work is an extremely valuable contribution to our understanding of
this beautiful and valuable genus. In current times, when researchers are
pressured to increase their publication list and H-indices, summarizing our
understanding about Pseudotsuga, as done here by the authors, is not likely to
happen. We are grateful that Denny and Dick maintained a commitment to its
completion during their retirement.
The compendium includes an invaluable collection of study descriptions and
results, supported by citations and references, many of which are only available
through paper copies. By synthesizing this information and making it available
on a digital, searchable platform, the authors have assured that this tremendous
body of knowledge will remain an extremely useful resource for future
generations of researchers and practitioners.
One impressive aspect of the book is its global coverage of Pseudotsuga. This
was only made possible by the authors’ global reputations and network of
colleagues around the world from whom they were able to obtain information,
and by their ability to read publications in numerous languages.
The book has an impressive scope and covers more than a century of
research results and management experience. It thus provides a historical
timeline that high- lights how challenges and information needs changed over
time and how research activities were centered around relevant issues.
Obviously, the authors relied heavily on their own work and experience, and
the literature coverage is especially detailed for the period from the 1960s to
1990s. Even as new knowledge becomes available, their groundbreaking
work as scientists and their synthesis efforts in writing this book will stand
the test of time. Their careers and work are a source of great pride for Oregon
State University, and this book will further strengthen the reputation of the
College of Forestry as a premier research and teaching institution.
On behalf of College of Forestry faculty, students, alumni, and also all
scientists and forest managers who stand to benefit from this book, the Dean and
Deans Emeriti thank Denny and Dick for this very significant contribution to the
field of forestry.

Thomas Maness, Dean, College of Forestry, 2012 –


present Hal Salwasser, Dean, College of Forestry, 2000 –
2012 George W. Brown, Dean, College of Forestry, 1990
– 2000
xvii
Preface
Pseudotsuga menziesii is distributed more widely outside its natural range than any
other species of American forest tree, with the exception of Pinus radiata. Its
successful introduction beyond its natural habitat into many parts of the temperate
regions of the northern and southern hemisphere is all the more remarkable because
of the early ignorance of, or disregard for, the importance of provenance.
The introduction of Douglas-fir went through various phases. Initially the species
was introduced through individual tree plantings in Europe and elsewhere around the
world. Successes were mostly dependent on the seed sources, and setbacks were due
to the occurrence of diseases, especially Rhabdocline pseudotsugae and Phaeocryptopus
gaeu- mannii. Over time, seed source problems and diseases were overcome through
genetic selection and silvicultural practices that allowed for wider establishment,
including monoculture stands. Social issues influencing such phases included trade
barriers, e.g., during war time, but more influential were the political discussions
about the intro- duction of non-native species. The proportion of Douglas-fir is often
limited because of concerns about its ecological impacts; for example Forest
Stewardship Certification standards in Germany limit non-native species to 20%
stocking in management units. Douglas-fir is nevertheless presently an accepted and
integral part of forest manage- ment in many countries because of its economic
importance, as well as its perceived reputation as a species that may be better able
to withstand some aspects of climate change, especially given its drought resistance.
But despite its unique characteristics and potential, Douglas-fir may also suffer from
potentially detrimental effects of a changing climate, such as warmer temperatures
that may not provide the critically important cooling period needed for flowering.
Indeed, a primary impetus for the development of this book came from our research
showing definitely negative effects of changing temperature that may negatively
impact Douglas-fir—making this “a species in peril.” This book discusses research
that demonstrates some of the unique aspects of Douglas-fir physiology. In it, we
provide an overview of research activities and find- ings that highlight unique
aspects of Douglas-fir physiology, genetics, and other related issues. The book
begins with the evolutionary history and distribution of Douglas-fir and provides a
detailed description of introductions of Douglas-fir to other countries, including
information about initial plantings, provenance trials, and genetic tree im-
provement activities. The sections about life history, drawn from extensive research
and teaching experiences, include detailed descriptions of flowering, seeds, root, and
seedling physiology, followed by sections about mycorrhizae and insects, diseases,
and other biotic factors.
This book is intended as a resource for everyone interested in understanding the
opportunities and challenges of managing Douglas-fir in a variety of regions and set-
tings. It provides information for historians and social scientists investigating
forestry trends; researchers, educators, and managers looking for detailed
information in areas such as genetics and regeneration practices; and all others
interested in the beautiful trees we call Douglas-fir.

xix
About the Authors
Denis P. Lavender is Professor Emeritus of Forest Science at Oregon State
University and former head of the Forest Science Department at the University of
British Columbia. He earned a bachelor’s degree in Forest Management from the
University of Washington, where he was named Outstanding Freshman Forestry,
1944; a mas- ter’s in forest science from Oregon State College in 1958; and a
doctorate in botany from Oregon State University in 1962.
After retiring as a professor in the OSU Forest Science Department in 1984,
Lavender went on to serve as head of the Forest Science Department at the
University of British Columbia from 1985 until 1992. While in Canada, Lavender
and several colleagues established the Silvicultural Institute of British Columbia.
He is also a Co-founder Forest Biology Workshop. Lavender has served as
president, executive secretary, and interim director of the Northwest Scientific
Association, which named him an honorary life member. He joined the Society
of American Foresters (SAF) as a member of the Willamette Chapter, where he
served as chairman, vice chairman, and secretary. Lavender is a member of the
Xi Sigma Pi Forestry Honor Society, the Phi Sigma Biological Sciences Honor
Society, and the Sigma Xi Scientific Research Society.
One noteworthy outcome of Lavender’s research in silviculture and plant
physiol- ogy over many decades is the creation of a methodology for the storage
and planting of seedlings that has increased conifer survival rate by 20%.

Richard K. Hermann is Professor Emeritus of Forest Resources at Oregon State


University. He earned a Diplom Forstwirt from Ludwig-Maximilians-Universität
München in 1951; and was awarded a full tuition scholarship to Yale University
in 1954, where he earned a master’s in forest soils science in 1956. At Oregon
State University, Hermann earned a doctorate in forest ecology in 1960. He was
awarded an honorary doctorate in silviculture from Georg-August-Universität
Göttingen, Germany, in 1979.
Hermann was a scientist in an exchange program between the U.S. National
Academy of Sciences and the Polish Academy of Science in 1971. In 1978, he
held a fellowship from the Institut National de la Recherche Agronomique
(INRA) of France at the Station d’amélioration des arbres forestiers, Centre
d’Orléans. He was guest professor in forestry at Georg-August-Universität
Göttingen in 1982 and 1989.
He has served as the leader of several working parties in the International
Union of Forest Research Organizations (IUFRO) Division 1 (1965–1981); as
Coordinator, Division 1 (1981–1986); and as a member of the IUFRO Executive
Board (1981–1986). Hermann was named Corresponding Member of
L’Accademia Italiana di Scienze Forestali (Italian Academy of Forest Sciences)
in 1982; an Oregon SAF Fellow by the Society of American Foresters in 1991;
and an Honorary Member of IUFRO in 1992. He is a member of SAF, the
Ecological Society of America, the NW Scientific Association, and the Sigma Xi
Scientific Research Society.
Hermann’s familiarity with both Douglas-fir in its native range and the
silvicultural practices in countries into which the species has been introduced
have enabled him to help his colleagues find solutions to problems they have
encountered in planting Douglas-fir, whether in the Pacific Northwest or around
the world.
xxi
Acknowledgements
The authors appreciate the support of the Oregon State University College of
Forestry and thank colleagues, staff, and students, especially in the Forest
Ecosystems and Society (Forest Science) Department. The authors also thank
Bonnie Avery at the OSU Libraries. Special thanks go to Jürgen Bauhus, Diane
Haase, Glenn Howe, Jürgen Huss, Scott Kolpat, Klaus Puettmann, and David
Shaw, who provided reviews.
The authors would like to acknowledge Martha Brookes, who edited early
drafts of the manuscript; Gretchen Bracher and Tristan Bahr, who created most of
the il- lustrations; student Alexa Pinckard, who typed much of the reference
section; and Bryan Bernart, who coordinated the project during interim stages,
edited chapters, and provided the cover photograph. Special thanks go to our
editor, Caryn M. Davis, who toiled tirelessly to enable the publication of this
book, and to Klaus Puettmann, who coordinated the review process.
Richard Hermann would like to acknowledge the contribution of Freya
Hermann, whose support made the completion of this book possible.

xxii
1. Evolutionary History
Richard K. Hermann

A
voluminous body of literature scattered and Mexico. Malyavkina
through a multitude of publications in many
languages covers the genus Pseudotsuga. The
tremendous proliferation of literature makes an
overall review of existing knowledge increasingly
difficult. Our book attempts to bring together what
is known about the genus Pseudotsuga.
The present distribution of the genus
Pseudotsuga is strongly discontinuous: it is
confined to western North America, Mexico, and
eastern Asia. The ge- nus includes eight to twelve
species of which two are indigenous to the United
States and Canada (Little 1979), four of
questionable status to Mexico (Martinez 1963),
four to mainland China, one to Taiwan (Li 1975),
and one to Japan (Ohwi et al. 1965).

The Genus Pseudotsuga


Douglas-fir belongs to the family Pinaceae, but is
unique among the Pinaceae in having a diploid
chromosome number of 26. All other species in
the Pinaceae, including big-cone Douglas-fir and
the Asian Douglas-firs, have a diploid
chromosome number of 24 (Doerksen and Ching
1972). As in most other conifers, Douglas-fir’s
nuclear genome (O’Brien et al. 1996) is large and
complex (3.7 x 1010 BP); chloroplast DNA is
inherited paternally (Neale et al. 1986) and
mitochondrial DNA is inherited ma- ternally
(Marshall and Neale 1992). Douglas-fir is
monoceous and has a mixed mating system
(selfing and out-crossing) but it mostly
outcrosses.
The fossil record suggests that intergeneric di-
vergence in the Pinaceae, on average, took place
135 million years ago in the Neocomian Epoch of
the Lower Cretaceous (Florin 1963). Fossil
remains of the genus Pseudotsuga from pre-
Cenozoic times have been reported from Russia
(1958) designated pollen from Lower have been advanced about the phylogenetic
Cretaceous deposits of the Eastern-Gobi position of Pseudotsuga in the family Pinaceae.
depression in eastern Mongolia as One theory holds that Pseudotsuga and Larix
Pseudotsuga punctata. Alvarez (1994) stated branched from a common lineage to Pinus
that 23 different kinds of fossil coniferous (Boureau 1938, Ferré and Gaussen 1945, Gaussen
pollen, including those of the genera Abies, 1955). The other hypothesis is that Pseudotsuga
Picea, Pinus, Larix, and Pseudotsuga, were descended from Larix which, in turn, supposedly
found in deposits from the Turonian Epoch descended from Pinus (Flous 1936, Campo-
(90–84 million years BP) of the Upper Duplan 1950, Gaussen 1966). The close rela- tion
Cretaceous in the Mexican State of Coahuila. between Pseudotsuga and Larix had already been
These assignments of fossil pollen to the pointed out in 1918 by Doyle, who called
genus Pseudotsuga, however, are not attention to the striking similarities between the
substantiated by megafossils. By the time two genera. They include anatomy of wood,
Pseudotsuga appears in the Tertiary fossil nonsaccate pollen, and structure of the female
record, it clearly resembles the modern genus. gametophyte.
Thus, we lack a record of most of the Results of immunological studies conducted
evolutionary history of the genus half a century later (Praeger et al. 1976, Price et
Pseudotsuga; all we have are hypotheses al. 1987, Price 1989) indicate that the close
based on phylogenetic studies. phylogenetic re- lation between Larix and
On the basis of morphological, anatomical, and Pseudotsuga proteins is consistent with
cytological studies, two principal hypotheses morphological evidence. Because

1
2 Douglas-fir: The Genus Pseudotsuga
the two lineages diverged, however, major pheno- of the Bering land bridge (Florin 1963). If some of the
typic changes arose between them. For example, fossil pollen of Upper Cretaceous age found in
Larix became deciduous but Pseudotsuga Coahuila indeed represent Pseudotsuga, that
remained evergreen.
Sziklai et al. (1987) used cytological
techniques to provide some insight into the
phylogeny of Pseudotsuga. Based on the
assumption that karyotype differentiation of
Pseudotsuga menziesii is the result of a
misdivision of a metacentric chromosome and the
production of stable telocentrics, they attempted
to reconstruct the 12 chromosomes thought to be
present in the ancestral Douglas-fir. This hypo-
thetical karyotype was used for numerical studies
of similarities among species in the genus, except
for the Asian Pseudotsuga brevifolia. Their
results of canonical variates analyses of
chromosome arm lengths suggest that
Pseudotsuga menziesii is more closely related to
Pseudotsuga macrocarpa than to any of the six
Asian species. In addition, their data show that the
karyotype of Pseudotsuga gausseni is the most
discrete of the Asian species and is also the most
similar of the Asian species to Pseudotsuga
menziesii. In an effort to shed more light on the
evolution- ary origins of the genus Pseudotsuga,
Strauss et al. (1990) used restriction fragment
analysis of chloro- plast, nuclear, and
mitochondrial DNA to study its phylogeny. Five
species of the genus, Pseudotsuga menziesii (Mirb.)
Franco, P. macrocarpa (Vasey) Mayr,
P. japonica (Shiras.) Beissn., P. wilsoniana Hayata,
P. sinensis Dode, and Larix occidentalis Nutt.,
were included in the study. These authors
concluded on the basis of their analysis that
Pseudotsuga evolved first in North America and
then migrated around the continental rim into
Asia. They found that ge- netic similarities
declined with migration distance around the
Pacific Rim.
The model for the origin of Pseudotsuga and
sub- sequent migration to Asia is consistent with
the fossil record. Fossils both mega- and micro- of
Pseudotsuga are represented in western North
America since about 50 million years ago, and in
Japan since about 15 million years ago (Hermann
1985). Mid-Tertiary fossils have been reported
(Wolfe 1969) from Homer, Alaska, near the
northernmost point of the migration route by way
too would support the hypothesis that the concluded that Pseudotsuga japonica is the Asian
genus evolved first in America. Alvarez species closest to the North American species, and
(1994) considered the fossil Pseudotsuga Sziklai et al. sug- gested that Pseudotsuga
reported from Coahuila to be the oldest gausseni is the Asian species most similar to
known in America, which, in his opinion, Pseudotsuga menziesii. The geographic
raised the possibility that the state of Coahuila
is the center of origin of the genus.
Strauss et al. (1990) noted that their data
suggest two principal hypotheses for
evolution within the genus. The first
hypothesis is that the bulk of the genus may
have been derived from a Pseudotsuga
macrocarpa-like ancestor, which then gave
rise to a P. menziesii-like line from which the
migrants to Asia were derived. The second
hypothesis would imply that a lineage
containing the future Asian migrant may have
split from the common ancestor of the genus
in North America before speciation of the
progenitors of present-day Pseudotsuga
menziesii and Pseudotsuga macrocarpa.
According to Strauss et al. (1990), for the first
hypothesis to be true, the 13-chromosome
karyotype of P. menziesii “would have to
have either (i) evolved after the split of the
Asian stock from it, (ii) remained
polymorphic within ancestral pre-
Pseudotsuga menziesii, eventu- ally become
fixed in P. menziesii and lost in the Asian line,
or (iii) been present but subsequently lost in
the Asian derivatives. Because of the
similarity of the Asian species’ karyotypes to
that of Pseudotsuga mac- rocarpa such a loss
or reversion is unlikely. Finally, and most
important, unless rates of evolution in these
lineages vary widely, an early split between
the North American species suggests that they
would have accumulated a greater genetic
distance between them in comparison to the
Asian species. This is not the case. The actin-
derived distance of 17.57 between
P. menziesii and P. macrocarpa was equivalent to that
between P. japonica and P. sinensis (18%)
(Strauss et al. 1990, Table 2).
Both the studies by Strauss et al. (1990)
and Sziklai et al. (1987) show that the North
American and Asian species form well-
differentiated monophyletic groups. Where
these authors differ is in their assign- ment of
the relation of two of the Asian species to
Pseudotsuga menziesii. Strauss et al.
Chapter 1. Evolutionary History 3
illustrate an advanced stage of subspecies evolution.
location of the range of P. gausseni, however, Its geographic races differ conspicuously and have
appears
to make such a relation unlikely.
The few studies about the evolutionary origins
of the genus Pseudotsuga have contributed to a
better knowledge of the taxonomic relationships
within the genus but provide no answers as to
when spe- ciation of modern Pseudotsuga
menziesii occurred. Morphological characteristics
of North American Douglas-fir little changed
since its appearance in the fossil record about 50
million years ago. The karyo- type of fossil
Douglas-fir is unknown and leaves us with the
question had Eocene Pseudotsuga sonomenis, the
ancestral Douglas-fir, already a haploid chromo-
some complement of 13, or evolved the n = 13
species some time later in the Tertiary from a
progenitor with a complement of n = 12.
Although the evolutionary history of the modern
Pseudotsuga menziesii remains obscure, the
suggestion that karyotype reduction and
fragmentation has been an evolutionary trend in
Pseudotsuga (Sziklai et al. 1987) would indicate
that Pseudotsuga menziesii is the most recent
species of the genus because P. menziesii has the
smallest chromosomes of all Pseudotsuga species.
The fossil record appears also to lend support
to
the assumption that P. menziesii is of rather recent
origin. In contrast to the Tertiary, Pseudotsuga is
often abundantly represented in Quaternary mega-
and microfossil assemblages. This would indicate
that Douglas-fir assumed its dominant role in the
forests of northwestern America in the
Quaternary. Wolfe (1969) emphasized that Early
Pleistocene assemblages still have low amounts of
Douglas-fir pollen whereas interglacial deposits
begin to show large quantities of pollen of this
genus. Therefore, he concluded that today’s
dominance of Pseudotsuga throughout much of
the conifer forest of western northwestern North
America was attained during the Middle or Late
Pleistocene. That change in the status of Douglas-
fir as a member of northwestern forests suggests
that a new species had evolved dur- ing the
repeated glacial cycles that could adapt to a wide
range of climates and site conditions.

Differentiation into varieties


Douglas-fir is one of several western conifers that
that the varieties evolved sometime in the Late
been recognized as the coastal variety, Pleistocene, during the Wisconsin glaciation
menziesii, and the interior variety, glauca, or (100,000–10,000 years BP) in glacial refugia, one
Rocky Mountain Douglas-fir. These races in the Pacific Coast region, and the other in the
appear to have evolved dur- ing repeated long southern Rocky Mountains.
periods of geographic isolation, alternating
with short flushes of gene exchange like the
present (Critchfield 1984).
Several hypotheses have been advanced
about when the two varieties of Douglas-fir
evolved. Axelrod (1980, pp. 72–73) proposed
that they had already originated during the
Tertiary. According to him, fossil evidence
indicates that the two varieties may have
evolved as early as the Oligocene and were
definitely in existence by the Miocene. The
Late- Oligocene flora of Creede, Colorado,
dated 27 mil- lion years BP, included
Douglas-fir with small cones characteristic of
the interior variety. Both varieties were
supposedly represented in floras of 13 million
years ago. Douglas-fir in the Purple Mountain
flora of western Nevada had small cones
(Axelrod 1980), but Douglas-fir in the Trout
Creek flora of southeast- ern Oregon had large
cones (Arnold 1935). The two varieties came
into contact with each other in central British
Columbia, northern Idaho, and north-central
and northeastern Washington about 7,000
years ago, after their Late-Quaternary
migrations from their respective glacial
refugia (Tsukada 1982).
Galoux (1956) suggested that both varieties de-
scended from a common ancestor in Mid-
Pliocene. In his opinion, the variety glauca
originated as the result of eastward migration
across the Great Basin and subsequent
adaptation to the environment of the Rocky
Mountains. He based his theory on the obser-
vation that the Pacific floral elements most
closely related to woody species in the Rocky
Mountains occur in the southern Cascades and
in the Sierra Nevada. An argument against
Galoux’s thesis is the statement by Sziklai et
al. (1987) that the smaller size of
chromosomes in the variety menziesii
suggests that it is a more recently derived
lineage than the variety glauca.
Halliday and Brown (1943), as well as
Heusser (1968), assumed a more recent
evolution of the two varieties. They proposed
4 Douglas-fir: The Genus Pseudotsuga

Perhaps the best available estimate of the time Intraspecific variation


since divergence between the coastal and interior The validity of recognizing two varieties of
variety was provided by Li and Adams (1989). Pseudotsuga menziesii has been questioned by
They suggested, based on their determination of Silen as recently as 1978 with the statement, “The
the av- erage genetic distance between 103 clinal nature of both morphological and chemical
populations distributed over the entire range of traits over the range of Douglas-fir and the
the species, that the two varieties have been in variability of types in a locality still raises doubts
existence for at least half a million years, and that
about the logic of varieties or subspecies within
they may have diverged during the Middle the species.” Nonetheless, Douglas-fir from the
Pleistocene, much later than the Pliocene but well western and east- ern parts of the species’ range
before the Wisconsin glaciation. Theshow sufficient differ- ences in their traits to make
differentiation of Douglas-fir into varieties may a case for a taxonomic division into a coastal and
have been influenced by large-scale climate interior variety.
change (Gugger and Sugita 2010). The differences in size and form of P.
menziesii within its range led, already
at the be- ginning of the 20th century,
to the real- ization that the division of
the species into a coastal and interior
CANADA
variety did not account adequately for
50° its variabili- ty. Frothingham (1909),
3 based on his de- tailed review of
Douglas-fir, divided its range into
five growing regions (Figure 1.1) in
an attempt to go beyond purely
O

45° taxonomic classification. Although he


arrived at the delineation of these re-
1
gions without the benefit of modern
4 bio- chemical and molecular genetic
tools,1 his scheme showed a
40°
remarkable degree of insight into the
2 regional differen- tiation of the
species. Ascherson and Graebner
(1913) acknowledged the exis- tence of
intermediates between coastal and
35°
5 interior populations and designated
these intermediates as variety caesia.
Region That taxonomic designation was
North Coast quickly accepted in Europe, in spite of
Sierra the lack of clearly distinguishable
Northern Rocky Mountain
Central Rocky Mountain
morphologi- cal characteristics.
Southern Rocky Mountain Schenck (1939) dis- tinguished a
MEXICO coastal (var. viridis) and two inland
varieties, var. caesia north, and var.
glauca south of lat 39° N. That

Figure 1.1 Frothingham’s (1990) division of the range of


Douglas-fir into five growing regions.
1. Molecular genetic markers are those derived from direct
analyses of genetic polymorphism in DNA sequences.
Biochemical markers are those derived from study of the chemical
products of gene expression, such as protein sequences or net
charges, and composition of secondary chemicals such as
terpenoids.
Chapter 1. Evolutionary History 5
latitudinal cline may exist, a suggestion previ- ously
geographic delineation of a northern and southern made by Sziklai (1969) and Chen et al. (1986).
inland variety comes close to the line of division
between a northern and southern inland chemi-
cal race that was made more than three decades
later by Zavarin and Snajberk (1973). Flöhr
(1958) with reference to Munns (1938) and
Schenck (1939) considered a relatively small area
in north-central Washington and south-central
British Columbia along the Frazer River south of
lat 50°30’ N as the area of transition between the
varieties menziesii and caesia. He assigned all
population west of that area of transition to the
variety menziesii and east of it to the variety
caesia, and those south of lat 41° N and east of
long 109° W to the variety glauca. The region
west of long 109° W and north of lat 27° N, up to
a line running northeast from lat 45° N to long
111° W, was designated as one that contained
populations of both variety glauca and caesia.
Morphological, anatomical, cytological, and
phys-
iological studies during the 1960s and 1970s
gradu- ally increased knowledge of intraspecific
variation patterns of Pseudotsuga menziesii. Allen
(1960b) and two of his students (Robinson 1963,
Dunlap 1964) demonstrated that the two varieties
of the species can be distinguished by the shape of
their seed. But they noted also that differences
become less distinct in the transitional zone from
the coastal to the drier interior region, pointing to
the existence of intermediate forms. Tusco (1963),
who had sampled 43 populations from coastal
British Columbia to the Porcupine Hills of
Alberta, has argued for the rec- ognition of
subspecific taxa in Pseudotsuga menziesii within
British Columbia based on the interpretation of
his data. Tusko’s interpretation, however, has
been challenged. Chen et al. (1986), in a study of
morphological variation of Douglas-fir in south-
western British Columbia, concluded that most of
it was within the chosen 46 populations. Hence,
these authors proposed that recognition of
subspecific taxa in this portion of the species’
range is inap- propriate. Similar results were
obtained by Scagel et al. (1987), who investigated
the variation of cone and seed morphology in 89
populations from west and east of the Cascade
Range in British Columbia, Washington, and
Oregon. They suggested, however, that a
provenances.
A growth chamber study involving 16 coastal Studies of wood properties of Douglas-fir
and 12 interior provenances (Nicholson 1963) (Griffin 1919, Miller 1961, Miller and Graham
had shown different growth responses to 1963, Bramhall 1966) also provided evidence of
short-day photoperi- ods between coastal and large variability within the range of the species.
interior provenances, and also between They demonstrated
provenances from the northern and southern
part of the interior variety’s range.
A study by Lavender and Overton (1972),
on the effects of a range of thermoperiods and
soil tempera- tures on the growth of Douglas-
fir seedlings raised from Vancouver Island,
western Washington, west- ern Oregon,
northern Montana, and New Mexico seed
sources, demonstrated distinct physiological
differences between the coastal and interior
prov- enances. Seedlings representing
populations from the northern and the central
portion of the coastal variety did not differ
conspicuously in their growth responses, but
seedlings from a southern Oregon seed
source exhibited a growth response that more
closely resembled the interior variety than the
coastal. In a study of intraspecific
variability of seed weight that used all 189
provenances from the 1966- 1969 IUFRO
seed collections, Birot (1972) distin- guished
a northern coastal group (British Columbia,
Washington, and northern Oregon); an east of
the Cascades group (east side of Cascades in
British Columbia, Washington, and Oregon);
a southern coastal group (southern Oregon
and California); and an interior group (New
Mexico, Arizona, Colorado, and Utah).
Except for the fact that the range of the
interior variety is not separated into a northern
and southern part, Birot’s grouping is very
similar to
that of Frothingham (1909).
Two studies (El-Lakany and Sziklai 1971,
1973) of variation in nuclear volume and
relative DNA content in 27 coastal and 25
interior provenances from the 1966-1969
IUFRO seed collection revealed that coastal
provenances have larger nuclear volumes and
greater relative amounts of DNA than the
inland provenances, and that a gradient exists
between coastal and interior populations. De
Vescovi and Sziklai (1975) were able to show
distinct differences. Values were significantly
lower for the inland than the coastal
6 Douglas-fir: The Genus Pseudotsuga
higher permeability of the wood of coastal Drew (1957) concluded that northern and
Douglas- fir than for interior Douglas-fir after south- ern wood property types could be
preservation treatments. Logs of eastern distinguished in the inland Douglas-fir. The
Washington and cen- tral-south British Columbia wood of the northern type is considerably more
origin showed inter- mediate degrees of dense and is stronger than that of the southern
permeability. Although wood of Douglas-fir from type. The distinction made by Drew between a
California does not reach the low permeability northern and a southern inland type was
values of variety glauca, Miller and Graham’s supported in the following decades by the
(1963) investigations indicate the existence of a results of terpene and allozyme analyses of
coastal California gradient with less permeable materials from the range of the variety glauca.
wood in the south and more permeable wood
Terpenes
towards the Oregon border. Sierra Nevada
material from Plumas and Nevada counties gave An important step towards more detailed
the lowest permeability values for logs from knowledge of intraspecific variation of Pseudotsuga
California. Differences in size and arrangement of menziesii was made with a series of chemo-
tracheids and position of the torus within the systematic studies. Von Rudloff (1972, 1973a,
bordered pit cavity between coastal and interior 1973b, 1975) demonstrated that analysis of leaf oil
populations and their intermediates are probably terpenes permits not only a clear- cut qualitative
responsible for the dif- ferences in permeability. and quantitative distinction between the two
recognized varieties of Douglas-fir but also
125°
120° 115°
a convenient quantitative descrip-
tion of interior and coastal
interme- diates. He emphasized
Kemano Prince George
British Columbia that this is possible because of
Alberta the presence of distinctly
Lonesome Lake Jasper different patterns of ter- pene
50° distribution and the relatively
Owl Creek small tree-to-tree variations
Nimpkish Clinton
within populations of the two
Tahsish Calgary varieties. Several chemical races
Mesachie Lake (Rudloff 1973a) differing
Vancouver
quantitatively in certain
IslandHope 50°
monoterpenes, appear to exist in
Snoqualmie Pass each variety (Figure 1.2).
In a later paper, von Rudloff
Elbe
Washington and Rehfeldt (1980) described
45° possible biosynthetic pathways,
Satus Pass
geographic variation, and
Montana
Santiam Pass inheri- tance of terpenes in
Douglas-fir from southwestern
Idaho
Canada and the northwestern
Northern Coastal Central Coastal Southern Coastal Coastal Intermediate
United Intermediate
States. Hypothetical
Olalla Interior Intermediate Rocky Mountain 45°
biosynthetic path- ways imply
Cave Junction Oregon
that geographic variation in 17
120° monoterpenes of the leaf oil can
be represented by three or four
s-fir distinguished by biogenetic pathways. Therefore,
penes. The broken line denotes the line of division between coastal and interior Douglas-fir as based on terpene patterns.
geographic variation between
the coastal and the inte- rior
variety can be
described by the
relative percentages
of β-pinene, the
terpinene-sabinene
group of
Chapter 1. Evolutionary History 7

terpenes, the camphene group, and perhaps limo- 1973).


nene. An abrupt transition between coastal and
interior varieties was found for terpenes of the
camphene group. An explanation for this kind of
transition is suggested by the results of progeny
tests with F1 intervarietal hybrids, which indicated
that high percentages of the camphene group are
controlled by a single dominant gene. Conversely,
relatively broad zones of introgression developing
from gradual changes in gene frequencies appear
to accompany geographic variation in β-pinene,
terpenes of the terpinene-sabinene group, and pos-
sibly limonene. Such a pattern is to be expected,
if quantitative inheritance is involved, again in-
dicated from results with F 1 intervarietal hybrids.
Von Rudloff and Rehfeldt (1980) cautioned that
the problem of constraint inherent in the use of
relative percentages imposes limitations on the
evaluation of genetic aspects. Nonetheless, their
findings with these intervarietal hybrids suggest
that the interme- diate terpene percentages found
in most trees in the zones of overlap are indeed a
measure of intermixing of the two varieties of
Douglas-fir.
Zavarin and Snajberk (1973, 1975, 1976) used
cortical monoterpenes to study geographic differ-
entiation of Douglas-fir throughout its range
except for Mexico. They distinguished four
chemical races (Figure 1.3): a coastal range in
western Oregon, western Washington, and western
British Columbia; a northern inland race in the
Rocky Mountains of Canada and the United States
north of the Snake River Basin (lat 42°30’ N); a
southern inland race in the United States south of
the Snake River Basin; and a Sierra Nevada race
in the central Sierra Nevada of California (Zavarin
and Snajberk 1973). They found that the northern
inland and the coastal race intergrade extensively
in central British Columbia, northeastern
Washington, and northern Idaho. Some
intergradation occurs also in the mountains of
central and east-central Oregon. California Sierra
Nevada populations are chemically different from
pure coast- al and interior populations, but show
closer affinity with the interior than the coastal
variety. Southern Oregon and coastal California
populations exhibit a chemically intermediate
status between variety menziesii and the Sierra
Nevada race (Figure 1.3; Zavarin and Snajberk
used tool than terpene analysis for assessments of
Zavarin et al. (1977) suggested that the the extent and pattern of genetic variation in
differen- tiation of Pseudotsuga menziesii into Douglas-fir. As gene markers, isozymes are
its southern and northern inland races was useful for describing
apparently brought about by the geologic
history of the Snake River basin. In the
Oligocene, an east-to-west trough, the
ancestor of the present day Snake River basin,
formed in central Idaho. This trough widened
and deepened in time, and thus separated from
each other the populations of coniferous
species growing in that area. The evolution of
the Sierra Nevada race is ex- plained by
Zavarin and Snajberk (1975) as a result of the
emergence of the Great Basin arid regions
during the Pliocene-Pleistocene epochs that
disjoined western Douglas-fir populations
from their eastern, inland counterparts.
Critchfield (1984) summarized the results of the
studies by Zavarin and Snajberk and von Rudloff
and Rehfeldt (1980) as follows,
The most conspicuous feature of terpene
variation in Douglas-fir is the uniformity of the
coastal race west of the crest of the Cascades and
north of the California border. Stands in this
region do not vary at all in a 3-carene, or
terpinolene. The uniformity ends abruptly at the
Siskiyou Mountains along the Oregon-California
border. The gradient connecting coastal and
interior races varies in width and steepness,
depending on the terpene. The sharpest
distinction is in the frequency of trees with high
levels of camphene-group terpenes, none in most
coastal stands and 100% throughout the north-
ern interior. Douglas-fir is highly variable in
terpenes throughout its California range with a
complex pattern of variation that may have
evolved over a long period under relatively stable
conditions. The interior region differs from the
Pacific region in lacking comparable
discontinuities.

By contrast, analysis of cortical terpenes from


sam- ples collected in the northern, central,
and southern parts of the range of big-cone
Douglas-fir indicated insignificant
intraspecific variability in that species
(Zavarin and Snajberk 1976). In addition, that
study did not provide any evidence of gene
exchange be- tween Pseudotsuga menziesii and
Pseudotsuga macro- carpa and corroborated
the work of Latling and Scora (1974) who had
shown good chemical separation between the
two species.
Enzymes
Electrophoresis has become a more widely
8 Douglas-fir: The Genus Pseudotsuga

140° 136° 132° 128° 124° 120° 116° 112° 108° 104°
52°
British Columbia

Approx. mid-point of Intermediacy 56°


Alberta

48°

Intergradation limits
Saskatchewan 52°

44° Coastal race


Washington Wisconsin Ice Front

48°

Montana

40°
Oregon Northern Inland race
Snake River Basin
Intergradation
Idaho 44°
Southern
Sierra Nevada race Wyoming Inland race
Utah
Nevada

36°
Colorado
40°

California
32°
New Mexico 36°

Arizona

0 200 400600 K
28°
32°
MEXICO Texas

124° 120°
116° 112° 108° 104°
Figure 1.3 Geographic differentiation of Douglas-fir based on cortical monoterpenes; extent of
Wisconsin glaciation colored dark (from Zavarin and Snajberk 1977).

genetic differentiation within and between taxa, 1977, Yeh and O’Malley 1980, Adams 1981,
and for verification of hybridity. Allozyme (allelic Hamrick et al. 1981) centering mainly on its
isozyme) analysis permits description of variation coastal variety demonstrated that the species has
patterns in terms of direct measures of genetic di- a great genetic diversity at enzyme loci. As in
versity (Adams 1981). Several allozyme studies of other conifers, most genic variation (95%)
Douglas-fir (Bergmann 1973, Muhs 1974, Yang et appears to be maintained within populations,
al. which may reflect the species’
Chapter 1. Evolutionary History 9

ecological amplitude, its breeding 130° 120° 110° 100°


sys- tem, and the lack of effective 55°
barriers to gene flow between
subpopulations (Yeh and O’Malley
1980).
Results of some of the allozyme
stud- ies indicate that northern and
southern populations differ markedly 50°

in variabil- ity at isozyme loci. Coastal Douglas-fir Northern Interior Douglas-fir Southern Interi
Critchfield (1984) considered
climatic conditions during the
Pleistocene to be responsible for
these differences because coastal and 45°
transitional populations of Douglas-
fir averaged 0.15 in heterozygosity CANADA
(Yeh and O’Malley 1980) and
interior populations 0.18 (Yeh 1981)
in glaci- ated British Columbia.
40°
Much higher levels were found by
other investiga- tors far south beyond
the boundary of the Wisconsin Ice
Front: 0.33 in coastal California
(Morris cited by Hamrick et al. 1981)
35° 81
and 0.26 in eastern Colorado 70 82
UNITED STATES
(Hamrick et al. 1981). Although
84
some weak clines in gene frequency
86
over environmental transects have 85
89 90
been re- ported among populations 91
30° 92
(Bergmann 1975, Mejnartowicz 87 93 94
96
1976, Yang et al. 97
1977, Yeh and O’Malley 1980), other 98
investigations (Merkle and Adams 101
1987; Moran and Adams 1989) did 102
25° 99
not show any association between
allozyme diversity and geographic
variables in intensive sampling of
100
Douglas-fir in southwestern Oregon,
a geographically restricted but
20°
environmentally diverse region of the
coastal variety.
A remarkable insight into the
geograpical patterns of genetic varia-
tion of Douglas-fir comes from an al-
lozyme study by Li and Adams
(1989), which used seeds from 104
MEXICO
sources dis- tributed over the entire
natural range of the species (Figure
1.4). The racial
patterns of allozyme variation found by these investigators conform closely to patterns determined
earlier from stud- ies of terpenes and Figure 1.4 Locations of seed from 104 sources used by Li and Adams (1989) in
their allozyme study.
quantitative traits,
10 Douglas-fir: The Genus Pseudotsuga
but they give a far more detailed picture of rather broad zones of transition except for terpenes
genetic variation in the species than heretofore
available. Cluster analysis based on genetic
distances between all pairwise combinations of
the 104 populations show that they clustered into
two groups corre- sponding to the recognized
coastal and interior va- rieties, except for one (no.
103) of the two Mexican provenances included in
the study (Figure 1.5).
The finding that Mexican population 103 from
Coahuila appears to differ genetically from all
other populations (Figure 1.5) may shed new light
on the taxonomic position of Mexican Pseudotsuga.
The large genetic distance (0.123) between
population 103 and the rest of the species implies
that the view that all of Mexican Douglas-fir is
part of variety glauca, held by North American
dendrologists (Harlow and Harrar 1969, Little
1979), probably needs revision. Incidentally,
Alvarez (1994) pointed out that popula- tion 103
came from the vicinity of General Cepeda, which
is near the locality where fossil Douglas-fir pollen
was found in an upper Cretaceous formation. The
interior populations separated into a north- ern
subgroup (British Columbia, Alberta, Idaho,
Montana, and northern Wyoming) and a southern
subgroup (central and southern Wyoming, Utah,
Colorado, Arizona, New Mexico, and Mexico) at
around lat 44° N.
Genetic structure of populations turned out to dif-
fer substantially among the three major
subdivisions. The coastal variety and northern
subgroup of the interior variety show considerable
genetic variation within populations but little
variation between them. By contrast, populations
in the southern interior subgroup are far more
genetically differentiated but have only about one
half the genetic diversity ob- served in coastal and
northern interior populations. Although range-
wide patterns of genetic varia- tion reported by
Li and Adams agree in general with those found
in previous studies of quanti- tative traits and
terpenes, they differ from them in some details.
Results of the allozyme analyses place zones of
transition between the coastal and interior variety
into south central British Columbia, north central
Washington, and central Oregon as did those of
earlier studies (Rudloff 1973a, Zavarin and
Snajberk 1973). The terpene analyses indicated
of the camphene group (Rudloff and Rehfeldt resistant to attack by the Cooley spruce gall adelgid
1980), but the data of Li and Adams suggest (Adleges [Gilletteella] cooleyi) than coastal
an abrupt transition regardless of whether the provenances farther north in Oregon, Washington,
two varieties are geographically separated as and British Columbia.
in central Oregon and north central
Washington, or are contiguous as in British
Columbia. Li and Adams speculated that their
inability to identify intermediate popula- tions
in British Columbia might reflect insufficient
sampling. They also considered the possibility
that gene flow between the varieties is not as
intensive in this region as was once assumed,
but that genes coding terpene variants may
have spread more rapidly because of selective
advantage.
The findings of Li and Adams (1989)
support separation of the interior variety into a
northern and southern subgroup as proposed
by Zavarin and Snajberk (1973) on the basis
of cortical terpene analysis. They placed the
break between the two subgroups at lat 42°30’
N but Li and Adams set it farther north at 44°
N. Li and Adams emphasized, however, that
separation between the subgroups seems
actually to be a gradual transition over at least
three to four degrees of latitude rather than an
abrupt change. The racial differentiation
shown by Li and Adams with allozyme
analysis differs in one major respect from that
found by Zavarin and Snajberk (1973). Li and
Adams did not identify a California Sierra
Nevada race with genetic affinity closer to the
southern race of the interior variety than to the
coastal variety. Their findings indicate a close
align- ment of the Sierra Nevada population
with those of the coastal variety. They pointed
out, however, that this difference in racial
patterning is not necessar- ily conflicting
because different traits may respond
differently to selection pressures. The
deviation in terpene composition of the Sierra
Nevada population from the rest of the
population of the coastal variety may be
related to differential selection pressures im-
posed by insects. As an example, Li and
Adams cite work by Stephan (1987) who
found that Douglas-fir provenances from
north coastal California, which is the
intergradation zone between the Sierra
Nevada and coastal terpene races identified by
Zavarin and Snajberk (1975), were more
Chapter 1. Evolutionary History 11

Genetic Distance General Location


0.14
0.12 0.10 0.08 0.06 0.04 0.02 0 No. State/Province Region
Coastal Variety
1 California Sierra Nevada
2 " "
4 " "
6 " Cascades
13 " Coast
45 British Columbia Vancouver Island
46 " "
5 California Cascades
14 " Coast
8 " Cascades
16 Oregon Coast
17 " "
18 " "
21 " Cascades
19 " "
20 " "
39 Washington W. Cascades
35 " "
10 California Coast
7" Cascades
34 Washington Coast
33 " "
36 " E. Cascades
48 British Columbia Cascades
9 California Cascades
30 Washington Coast
12 California "
31 Washington "
15 California "
44 British Columbia Vancouver Island
11 California Coast
32 Washington Coast
38 " E. Cascades
37 " W. Cascades
47 British Columbia Coast
49 " "
52 " "
50 " "
51 " "
3 California Sierra Nevada
22 Oregon Central
23 " "

Interior Variety-
Northern Subgroup
24 " "
27 " Eastern
25 " Contral
26 " Eastern
58 British Columbia Rockies
28 Oregon Eastern
29 Washington Southeastern
41 " Northeastern
43 " "
65 Idaho
71 Montana
72 "
74 "
75 "
68 Idaho
78 Montana
80 Wyoming Northern
63 British Columbia Rockies
60 " "
42 Washington Northeastern
64 Alberta
66 Idaho
77 Montana
67 Idaho
79 Montana
88 Colorado Northern
76 Montana
69 Idaho
81 Wyoming Northern
40 Washington E. Cascades
56 British Columbia Rockies
53 " "
73 Montana
54 British Columbia Rockies
55 " "
57 " "
59 " "
62 " "
61 " "

Interior Variety-
Southern Subgroup
82 Wyoming Central
70 Idaho Southeastern
84 Utah
87 "
83 Wyoming Central
86 Utah
85 "
93 Colorado Southern
89 " Central
90 " "
96 " Soutern
92 " Central
100 New Mexico
91 Colorado Central
97 New Mexico
95 Colorado Southern
94 " "
98 New Mexico
102 Arizona
101 "
99 New Mexico
104 Mexico
103 Mexico Coahuila

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0


Genetic Distance
Figure 1.5 Genetic distances between provenances used for the range-wide allozyme study (from Li and Adams 1989, modified).
12 Random amplified polymorphic DNA
Douglas-fir: The Genus Pseudotsuga That transect, which crosses the putative transition

The use of random amplified polymorphic DNA zone between the two varieties in central Oregon,
(RAPDs) is the most recent technique applied to is essentially the same used earlier by Sorensen
the investigation of genetic variation in Douglas- (1979) for a common garden study. Seeds of the
fir. Aagaard et al. (1995) studied racial differen- southern interior race of the interior variety
tiation and genetic variability between and within stemmed from three locations in Utah, two in
the coastal, north interior, and south interior races Arizona, and four in New Mexico. For a
of Douglas-fir by means of RAPD and allozyme comparison of genetic differentia- tion and
markers. Seed samples of the coastal variety and diversity between RAPDs and allozymes,
the north interior race of the interior variety came Aagaard et al. (1995) matched their 29 seed
from 20 locations along an east-west transect from sources with 31 populations from Li and Adams’
just west of the Cascade summit to eastern Idaho. (1989) range-wide allozyme study of Douglas-fir
(Figure 1.6) and re-analyzed data from these 31
populations.
125° 120° 115° 110° 105°

CANADA Coastal North Interior South Interior

50°

Washington
45° 65 73
18 21 24 28 66 75 Montana
1720 78
27
22 25
16 19 2326 45°
67 69
Oregon 68 Idaho
Wyoming
40°
84
Nevada 85 86

40°

Colorado
87 Utah
35° California

101
98
102
New Mexico 35°
Arizona
99
30°
100

Texas
MEXICO

120°
115° 110° 105°

0 200 400 600 Figure 1.6 The 29 seed sources matched with 31 populations
K from Li and Adams (1989)range-wide allozyme study.
Chapter 1. Evolutionary History 13

Table 1.1 Racial differentiation, genetic distances, and genic and allelic diversity based on RAPDs and allozymes between or within races of
the coastal and interior varieties of Douglas-fir. Standard errors are calculated for GST from variance among locus-specific estimates.

RAPDs Allozymes
Differentiation (GST) between races
Coastal × north interior × south interior 0.73 ± 0.09 0.26 ± 0.03
Coastal × north interior 0.76 ± 0.12 0.21 ± 0.03
Coastal × south interior 0.82 ± 0.09 0.36 ± 0.04
North interior × south interior 0.52 ± 0.11 0.19 ± 0.03
Gene diversity (HS) calculated within races
Coastal 0.07 ± 0.04 0.21 ± 0.04
North interior 0.15 ± 0.05 0.17 ± 0.05
South interior 0.18 ± 0.05 0.09 ± 0.04
Mean gene diversity (HT) averaged over races 0.13 ± 0.03 0.16 ± 0.03
Total gene diversity 0.49 ± 0.05 0.21 ± 0.05

Results of the RAPD study revealed striking 130° 120° 110° 100°
contrasts between estimates of racial
differentia- tion and diversity with RAPD and
allozyme mark- ers (Table 1.1). Differentiation CANADA
between races of Douglas-fir based on RAPD 9
50° 8 1 55°
analysis accounted for
more than 70% of the total diversity observed (GST =
0.73; Aargard 1995). Conversely, of the total
genetic variability found for allozyme data, less 45° 7 50°
than 30% (GST = 0.26) was attributed to racial
6
differentiation. 3
Furthermore, allelic and genic diversities between UNITED STATES
45°
races were inverted for RAPDs relative to that for 40° 5 2
allozymes. Gene diversity for RAPDs was highest 4
in the south interior race (HS = 0.18) and lowest in
the coastal race, but the reverse was true for 40°
allozymes (Table 1.1). The observed number of 35°
alleles per locus paralleled these trends, with the
highest values in the south interior race (1.44) and 35°
the lowest in the coastal race (1.25). 30°
Results of the study of RAPD phenotypes
pointed to a sharp boundary between the coastal
30°
and north interior races. The break occurred
between the lowest elevation Santiam population 25°
(915 m; lat 44°25’ N, long 121°38’ W) and the
Grizzly population (1,555 m; lat 44°26’ N, long 25°
120°57’ W), which are separated by a mere 55 20°
km. That finding supports the sugges- tion by Li MEXICO
and Adams (1989) of an abrupt transition between 20°
the coastal and northern interior races, although
their allozyme study indicated the pos- sibility of 115° 110° 105° 100°
a narrow transition zone between these races. The
abrupt transition between the coastal and interior
varieties found by Zavarin and Snajberk
(1973) for terpenes of the camphene group suggesting
agrees most closely with the RAPD data by
Figure 1.7 Ecotypes of Douglas-fir based on
geographic variation of allelic structures (from
Klumpp 1999).
14 Douglas-fir: The Genus Pseudotsuga
a distinct racial boundary in nearby populations. The results of an isozyme study by Klumpp
Aagaard et al. (1995) pointed out that intervarietal (1999), based on commercial seed from 27
hybrids of Douglas-fir are known to readily occur sources within the natural range of Pseudotsuga
in zones of contact, and gene flow through pollen menziesii and buds from 11 populations in German
and seed dispersal is extensive in conifers. From provenance trials, are essentially in agreement
that information, they inferred that the close with those of Li and Adams (1989). Klumpp
proximity of coastal and interior varieties in recognized three major forms of the species,
central Oregon may have come about in recent coastal, northern, and southern inland Douglas-fir.
geologic time. Aagaard et al. (1995) summed up But he went one step further than Li and Adams
their comparison of RAPD and allozyme markers (1989) and distinguished nine ecotypes (Figure
as follows: “RAPDs appear to provide greater 1.7) on the basis of the geographic variation of
sensitivity than allozymes for the detection of allelic structures. Klumpp, however, considered
genetic differentiation, at least for long- isolated these additional subdivisions as provisional in
gene pools such as races of Douglas-fir.” view of his limited sample size.
2. Natural Range
Richard K. Hermann

D
ouglas-fir (Pseudotsuga menziesii (Mirb.) botanist Franco (1954) proposed Pseudotsuga
Franco growing on the moist Pacific slopes from menziesii (Mirb.) Franco as the valid scientific
British Columbia to California, commonly name that has won universal acceptance.
referred to as the Douglas-fir region, is the com-
mercially most important tree species in western Varieties of Pseudotsuga menziesii
North America. In its drier interior range the The British botanist Murray was the first to
species has been rapidly gaining in commercial recognize two geographically distinct groups of
importance since the middle of the 20th century Douglas-fir in 1869 (Little 1952). They are now
(Hermann and Lavender 1990). Douglas-fir has recognized in North America as the coastal
been a major com- ponent of the forests of variety (Pseudotsuga menzie-
western North America since the middle-
Pleistocene (Hermann 1985). Although the fossil
record indicates that the native range of the
species has never extended beyond western North
America, it has been successfully introduced into
many regions with a temperate climate since the
mid-19th century (Hermann and Lavender 1999).
The introduction of Douglas-fir into temperate
zones of both hemispheres has been expanding in
the 20th century because appropriate provenances
often outgrow the native conifers on suitable sites
(Hermann 1987).

History
Douglas-fir was discovered by Archibald Menzies
(1754-1842) on the west coast of Vancouver
Island. Menzies, who accompanied Captain
Vancouver on journey to northwest America as
ship surgeon on board the Discovery, did not
mention the tree in his journal (Menzies 1923).
He brought back, however, a specimen of
Douglas-fir which the British botanist Lambert
had described in 1803 as Pinus taxifolia. For his Figure 2.1 Lambert’s (1832) botanical illustration of Douglas-fir.
original description, Lambert (1803) did not have
cones; however, he included them in an illustra-
tion for the 1832 edition (Figure 2.1). Douglas-fir 1. The International Code of Botanical Nomenclature specifies that
underwent several changes in its scientific name the variety that includes the type specimen of the species must bear
the same epithet as the species without citation of author (Little
in the next 150 years (Hermann 1982). The 1952).
Portuguese
15
16 Douglas-fir: The Genus Pseudotsuga
1
sii var. menziesii) and interior variety Coastal Douglas-fir commonly reaches ages of
(Pseudotsuga menziesii var. glauca Beissn. 750 years (Franklin and Dyrness 1973) but may
(Franco) (Little 1979). The two varieties live longer. The oldest coastal Douglas-fir on
intergrade in areas of contact from the northern record was found near Mount Vernon,
half of Oregon northward into central British Washington. The age count was made on a section
Columbia (Rudloff 1972, Sorensen 1979). In about 12 m above the stump. The age at that point
contrast RAPD markers amplified from mitochon- was 1,375 years. Since at least 25 years were
drial DNA showed a rather abrupt genetic required to reach the height at which the age count
disconti- nuity in this area (Aagard et al. 1995). was made, the tree must have been over 1,400
The varieties are morphologically, physiologically, years old when cut in 1913 (McArdle and Meyer
and chemically distinct (Silen 1978). Needles of 1949). On the best sites, coastal Douglas-fir can
the coastal variety are green; presumably that is become huge; the tallest on record was cut in
why the variety was referred to in the older 1895 in the Capillano Valley near Vancouver,
European literature as var. viridis (Latin for British Columbia. It measured 127.1 m from
green). Needles of the interior va- riety are ground level to the tip of the leading shoot. That
blueish. Either color of foliage is occasion- ally is, 15.8 m taller than the tallest then-known
displayed by both varieties. Structure of their Sequoia sempervirens in California (Edlin 1965).
cones differs also. Cones of coastal Douglas-fir A 1,022-year-old coastal Douglas-fir cut in 1924
have straight, flat bracts while those of the interior near Mineral, Washington was 119.3 m tall, with a
vari- ety have exserted and strongly reflected diam- eter of 4.9 m and a volume of 249 m3
bracts. In general, coastal Douglas-fir lives longer (O’Brian 1994). The Mineral tree was 9 m taller
and reaches greater height, diameter, and volume than the 110 m-tall coast redwood at Dyerville,
than interior Douglas-fir (Table 2.1). The data in California (Bronaugh 1992). Interior Douglas-fir
Table 2.1 were compiled from second- and third- rarely grows older than about 400 years
growth stands and, thus, do not show dimensions (Frothingham 1909). The oldest liv- ing interior
known to have been attained by old growth Douglas-fir ever reported was found on
Douglas-fir.

Table 2.1 Individual tree characteristics for Douglas-fir in the western United States (modified from Van Hooser et al. 1991).

Diameter at breast height (cm) Height (m) Volume (m3)


Mean Max Mean Max Mean Max
Interior Douglas-fir region/ Northern
Rocky Mountains
Idaho 34.5 204.7 19.5 61.0 1.08 24.5
Montana 28.7 120.7 15.8 54.3 0.62 14.4
Wyoming 28.7 107.2 13.1 32.3 0.65 6.4
Interior Douglas-fir region/
Southern Rocky Mountains
Arizona 18.5 38.6 10.4 21.6 0.14 0.7
Colorado 27.4 106.4 12.5 38.7 0.59 6.3
New Mexico 23.6 84.8 11.9 32.0 0.51 4.7
Utah 31.5 113.5 14.9 38.1 0.82 8.5
Pine subregion
Eastern Oregon 25.4 129.0 16.7 51.5 0.42 11.7
Eastern Washington 25.4 192.8 18.3 70.7 0.48 36.9
Coastal Douglas-fir region
Western Oregon 27.2 195.6 20.1 78.9 0.71 43.6
Western Washington 27.7 192.5 22.6 78.3 0.79 45.2
California 28.9 186.0 18.9 71.3 0.96 48.0
Chapter 2. Natural Range 17

the Bandera lava flows in the 60°


El Malpais National
Monument, New Mexico
(Swetnam and Brown 1992).
It has a pith date of CE 1062.
Compared to coastal Douglas- 55°
fir, the interior variety tends to
be slower-growing, more cold
hardy, more drought hardy,
50°
more shade tolerant, and more
susceptible to Swiss needle cast
(Phaeocryptopus gaeumannii).
45°
Range
Early botanical explorers of
the American west, such as 40°
David Douglas, were already
aware of the extensive range of
Douglas- fir. Hooker (1838) 35°
wrote in his Flora Borealis-
Americana: “Mr. Douglas
30°
observes that the prin- cipal
part of the gloomy forests of
Northwest America, in the val-
25°
leys of the Rocky Mountains,
and throughout the interior
skirting those mountains, is 20°
composed of this species.”
Nearly a century would pass,
15°
however, between the
discovery of Douglas-fir by 125° 120° 115° 110° 105° 100° 95° 90° 85° 80°

Archibald Menzies in 1792 at


Nootka Sound and the first
comprehensive description of
its
natural range by Sargent Figure 2.2 Natural range of Douglas-fir (from Little 1971).
(1884).
Frothingham (1909) published
the first distribution map of Pseudotsuga Mountains of Canada and the United States into the
menziesii, including both the coastal and interior mountains of central Mexico over a distance of
varieties. Subsequently, this map was revised by nearly 4,500 km. The western half is
Sudworth (1918), Munns (1938), Fowells (1965),
and Little (1971). The latitudinal range of
Douglas-fir (Figure 2.2) is the greatest of any
commercial conifer of western North America.
The range resembles an inverted V with uneven
sides. From the apex in central British Columbia,
the western half extends along the Pacific
mountain ranges into California for about 2,200
km. The eastern half stretches along the Rocky
considered as representing the range of
Pseudotsuga menziesii var. menziesii, and the
eastern half the range of Pseudotsuga
menziesii var. glauca.
Sargent’s statement, “The line which
marks the northern limits of its distribution as
now known is curiously irregular” is still
valid (Figure 2.3). The most northerly record
of living Douglas-fir in the Prince George
Forest District is Tudyah Lake, lat 55°05’ N,
long 123°00’ W. The species reaches farther
northward, to 55°30’ N at Babina Lake and
Takla Lake (Garman 1963, Revel 1966,
unpublished paper) in the eastern part of
Prince Rupert Forest District. From there, its
boundary runs in a general south-
southeasterly direction to the headwaters
18 Douglas-fir: The Genus Pseudotsuga
135° 130° 135° Within its range Douglas fir is able to grow to
altitudes as great or greater than western red cedar,
and reaches within a few hundred feet of western
hemlock. It would be reasonable to expect that the
upper altitudinal limits of these species are
British Columbia primarily controlled by factors as- sociated with
decreasing temperatures occurring with increasing
altitude. It would also be reasonable to expect that the
northern latitudinal limits would come under the
control of temperature; and therefore, the order of
species limits should be the same latitudinally as
altitu- dinally. However, this is not the case. The
50° northern limit of Douglas fir stops almost 480 km
south of western red cedar and nearly 1,120 km
south of western hemlock. These anomalies support
the view that northward migra- tion of Douglas fir
has been halted by a low fire hazard barrier,
whereas red cedar and western hemlock have
successfully colonized their entire potential range
by virtue of their capacity to reproduce in shade as
well as
Figure 2.3 Line designates the northern limit of Douglas-fir in 2. R. L. Schmidt, British Columbia Forest Service.
British Columbia (Schmidt 1965, unpublished paper). 3. Philip G. Haddock, University of British Columbia.

of the Nazko River, then in a west-southwesterly


direction to the headwaters of the Klinaklini
River. From here, the line follows a generally
northwesterly direction to Bella Coola, and finally
turns south, pro- ceeding to the coast. From Bella
Coola, a small tongue extends northward and then
divides, Douglas-fir reaching up the Dean River
Valley in an easterly direction, and the other half
extending in a north- westerly direction into the
vicinity of Kemano (lat 53°30’ N). The presence
of Douglas-fir on the Skeena River (lat 54°20’ N)
is mentioned by several authors (Sargent 1898,
Sudworth 1908, Frothingham 1909) and is shown
as an “insular” occurrence on all range maps of
Douglas-fir published, except that by Krajina et
al. (1982). The older maps are in error. Neither
R.L. Schmidt2 (personal communication) nor P.G.
Haddock3 (personal communication) could
confirm an outpost of Douglas-fir on the Skeena
River.
An interesting point was made by Schmidt
(1960) about the northern limit of Douglas-fir in
coastal British Columbia. He argues that the
present north- ern boundary of the range of
coastal Douglas-fir reflects low fire frequency
rather than climatic con- trol as postulated by
Henry and Flood (1920, p. 71). Schmidt (1960)
saw supporting evidence for his claim in a
comparison of the altitudinal and latitudinal limits
of Douglas-fir and its principal associates:
on denuded areas. (Schmidt 1960, pp. 156–57)

Little’s (1971) map indicates a


discontinuity of the range of Douglas-fir in
the area of the Seymour- Neechanze Rivers.
The inventory data of the British Columbia
Forest Service, however, suggest continu- ous
distribution of Douglas-fir along these
drainages. Little’s map further shows absence
of Douglas-fir in a continuous strip in the
Coast Range of British Columbia extending
southeastward from about lat 53° N to slightly
below lat 50°N. Actually, the dis- tribution of
Douglas-fir is continuous through the Coast
Range along several corridors. North, east,
and south of Haylmore is a large enclave
without Douglas-fir. The boundaries of
Douglas-fir in the area delineated by long
122°30’ W and 124°10’ W and by lat 51° N
and 51°50’ N are uncertain (R.L. Schmidt,
personal communication). The species is
absent from the northern tip of Vancouver
Island and the coastal archipelago. Douglas-fir
ranges continuously through western
Washington and western Oregon. The eastern
slopes of the Cascades form the eastern
boundary. Sudworth’s (1908) report that the
species is absent from the east side of the
Cascade Range south of lat 45° N is incorrect.
Douglas-fir continues along the upper east-
side slopes and stops short of the Oregon-
California line (Franklin and Dyrness 1973).
The range forks in northern California between
lat 40° N and 41° N.
In the West, Douglas-fir extends along the
California Coast Ranges into the northern end
of Sonoma County. From here on southward,
the range becomes discontinuous. The largest
discontinuous area of the species is in the
Santa Cruz Mountains
Chapter 2. Natural Range 19
barely
between San Francisco and Monterey Bay. A
stand along Salmon Creek south of Los Burros in
the Santa Lucia Mountains was the most southerly
known (Langenheim and Durham 1963) until
1964. That year Griffin reported the discovery of
a grove of Pseudotsuga menziesii in a deep,
narrow canyon in the Purisima Hills (lat 34°44’
N, long 120°25’ W) near Lompoc, 145 km
southeast of Salmon Creek. He considered this
stand a Pleistocene relict but did not dismiss
entirely the possibility that the trees had been
planted. Analysis of cortical monoterpenes
(Zavarin and Snajberk 1975) from trees of the
Lompoc stand, however, indicated that the stand
belongs to their Sierra Nevada chemical race and
fits exceedingly well into the south-to-north
chemical gradient along the coast, which lends
strong support to the idea that the Lompoc stand
represents a Pleistocene relict.
The eastern half of the range in California ex-
tends continuously in a south-southeasterly direc-
tion through the Sierra Nevada into the southern
part of Yosemite National Park. The southern limit
in the Sierra Nevada is at Big Creek (lat 37°11’ N,
long 119°15’ W; elevation 1490 m) about halfway
between Yosemite and Kings Canyon national
park’s boundaries (Zavarin and Snajberk 1975).
The stand at Big Creek is limited in extent and
isolated from the main distributional range of the
species. The species is found on the east side of
the Sierra Nevada, al- though it is extremely rare
there (Frothingham 1909).
Beginning again at the northernmost
occurrence of Douglas-fir is the starting point for
the descrip- tion of the eastern half of its range.
The northern boundary runs from Takla Lake east
southeast to McLeod. There, the boundary of
Douglas-fir turns in a southeasterly direction
representing the eastern limit along the
Continental Divide to the Canada- United States
line.
Reports on the northern limit of the species be-
tween Takla and McLeod are at variance. Little’s
map shows an outlier, which according to R.L.
Schmidt (personal communication) is not separate
from the main distribution shown by the range
map of Douglas-fir in British Columbia (Figure
2.4). Douglas-fir is reliably reported from Tudyah
Lake (P.C. Haddock, personal communication),
Miette River valleys around Jasper town. Small
to the Peace River. The boundary line is groves and isolated trees occur sporadically in
difficult to draw in that area because small the foothills and upper North Saskatchewan
groves of Douglas- fir and individual trees River Valley. The northern recorded limit of
that survived fires are scat- tered in the region Douglas-fir in Alberta is near Brule Lake
north of Prince George to about lat 55° N. (53°15’ N, 117°50’ W) (Stringer and LaRoi
None of the reports cited above confirms the 1970).
statement by Halliday and Brown (1943) that The range of Douglas-fir is fairly continuous
the northern limit of Douglas-fir is near the through northern Idaho, western Montana, and
Finlay Forks on the 56th parallel. northwestern Wyoming. A large area containing
Another mapping problem is posed by Douglas-fir extends southwestward from central
the grass- land areas in central British Idaho into the Wallowa and Blue Mountains of
Columbia. Douglas-fir is scattered Oregon. Several outliers of Douglas-fir are
throughout these areas, but Little (1971) present in Alberta, Montana, and Wyoming, the
mapped these areas as not containing largest in the Bighorn Mountains. From southern
Douglas- fir. Contrary to older descriptions, Idaho south- ward through the mountains of Utah,
such as that of Sudworth (1908), the species Colorado, New Mexico, and Arizona, the
is represented in the Caribou Range and distribution becomes discontinuous. “The species
high elevations in the Gold and Selkirk is absent or rare in the dry interior basins, and on
Mountains (Figure 2.4; J. Revel,4 P.G. the semiarid plateaus and minor mountain spurs
Haddock, personal communication). lying between the principal ranges, especially
Douglas- fir is abundant in the Porcupine toward the southern and eastern limits of its
Hills, and Bow, Kananaskis, and Oldman range” (Frothingham 1909). Distribution in the
River valleys south and west of Calgary Rocky Mountain region has now been more ac-
(Moss 1944, Bird and Hong 1969). curately mapped, although minor revisions may
Elsewhere in Alberta, Douglas-fir forests still become necessary. Thus, Critchfield and
are com- mon only in the Athabasca and Allenbaugh
over the lat 55° N line and closely adjacent to the
Hart Highway leading north from Prince 4. J. Revel, British Columbia Forest Service.
George
20 Douglas-fir: The Genus Pseudotsuga

56°

British Columbia 58°


Alaska

Alberta

54°
56°

52° Prince George


54°

Interior Plateau
Bella Coola
Williams Lake
50°
52°

Kamloops Golden

48°
Vancouver 50°
Nelson
Distribution of Douglas-fir
0 200 K
Washington Idaho
Montana
128° 124° 120° 118°

Figure 2.4 Range map of Douglas-fir in British Columbia (from Revel 1966).

(1969) pointed out that Douglas-fir extends north- Durango, and Zacatecas. In the
westward from the Wasatch Range into the Raft
River (Preece 1950) and Albion Mountains. The
species stops short of the South Hills of southern
Idaho but is mapped there by Johnson (1961) and
Fowells (1965). Critchfield and Allenbaugh
(1969) also doubted the occurrence of Douglas-fir
in the mountains of Elko County in northeastern
Nevada as reported by Little (1956). In addition,
they recorded two sizable but previously
unreported outliers of Douglas-fir in the Owyhee
Range of southwestern Idaho.
The most southerly occurrence of Douglas-
fir in the United States is in extreme western
Texas. The tree is present in the Guadalupe
Mountains which extend from southeastern
Arizona into Texas, in the Sierra Vieja, and
farthest south in the Chisos Mountains (lat
29°13’ N).
In Mexico, Douglas-fir extends
discontinuously along the Sierra Madre
Occidental and is present in Sonora, Chihuahua,
Sierra Madre Oriental, the species is reported that Douglas-fir in Mexico does not form large
from Coahuila, Nuevo Leon, Tamaulipas, stands (Blanco 1941) and usu- ally is a minor
Hidalgo, and Puebla (Figure 2.5). The component of the forest. Until 1994, lat 19° N in
description of the range of Douglas-fir in Puebla was the known southern limit of Douglas-
Mexico by Martinez (1949) agrees in general fir. The discovery of Douglas-fir in the state of
with Little’s (1971) map, except that Little Oaxaca has extended the known distribution of
does not show the species to be present in the genus Pseudotsuga to 16°22’ N, 96°06’ W,
Tamaulipas. Martinez’s account that it “then 110 km southeast of the town of Oaxaca. The
extends from Tamaulipas toward the south locations of three isolated stands, each of about 2–
through the cen- tral and eastern region to the 3 ha, have been reported so far from Oaxaca
north of Puebla” could be interpreted to mean (Debreczy and Racz 1995).
Chapter 2. Natural Range 21

Altitudinal distribution Mount Hood, Douglas-fir extends to an altitude of


Altitudinal distribution of both varieties of Douglas- 2,195 m according to U.S. Forest Service
fir increases from north to south reflecting climatic inventory records. In the southern Oregon
control over distribution of the species. The Cascades and in the Sierra Nevada, the altitudinal
principal limiting factors are temperature in the range is gener- ally between 610 m and 1,829 m.
north, and moisture in the south of the range. As a In river valleys and canyon bottoms, the species
general rule, the interior variety grows at may occasionally descend to elevations of 244 to
considerably higher al- titudes than the coastal 274 m. Below 610 m, however, Douglas-fir is
variety at the corresponding latitude. For example, infrequent and is scrubby in appearance. Near the
at lat 45° N, Douglas-fir oc- curs up to an elevation southern limit of its range, in the Sierra Nevada,
of about 1,220 m in the Coast Ranges, and up to the species will grow to eleva- tions of 2,286 m
2,440 m in the Rocky Mountains. Whether this according to Frothingham (1909). Sudworth
distribution is a genetic adaptation or reflects (1908) listed the highest altitude at which
climatic differences is unclear (Silen 1978). Douglas-fir occurs in the Sierra Nevada as 2,225
Altitudinal limit for Douglas-fir in central British m at Glacier Point in Yosemite National Park. In
Columbia is about 760 m but rises to 1,250 m on the Santa Cruz and Santa Lucia Mountains, the
Vancouver Island (Heusser 1960). In Washington lower altitudinal limit for Douglas-fir is 762 m.
and Oregon, the species ranges from sea level to Elevation of the southernmost stand of the coastal
1,524 m, although locally it may occur higher. On variety in the Purisima Hills is 213 m (Griffin
1964).
The inland variety grows at altitudes from 549 to
2,438 m in the northern part of its range
(Kirkwood 1922). In Jasper and Banff national
parks, the upper altitudinal limit for the species is
at 1,372 m (Heusser

Baja
California

Sonora 100°
UNITED STATES
Chihuahua

MEXICO Coahuila de
Zaragoza

Baja San Luis


Sinaloa
California
Sur
Durango
Zacatecas
e e
N v L ó
er
u o n Tamaulipas
Aguascalientes Potosí Querétaro
Nayarit
Guanajuato
Tlaxcala Yucatán
Jalisco Hidalgo
México Quintana
D.F. Roo
Colima Michoacán Morelos
Puebla Veracruz Campeche
Tabasco
Guerrero BELIZE
Chiapas
16° Oaxaca

0 200 400 GUATEMALA


km
0 200 200
mi

Figure 2.5 Range of Douglas-fir in Mexico (modified from Martinez 1963).


22 Douglas-fir: The Genus Pseudotsuga
5. Frederick C. Hall, USDA Forest Service, Region 6.
1968) and increases gradually to 2,438 m in
6. John R. Jones, USDA Forest Service, Rocky Mountain Forest and
Montana, Idaho, and northern Wyoming. Range Experiment Station.
According to Hall5 (personal communication),
the lower altitudinal limit of Douglas-fir in the
Blue Mountains of northeastern Oregon is at 518
m, and the upper limit is at 2,134 m. In the central
Rocky Mountains the species is mostly found at
elevations between 1,830 m and 2,592 m (Bates
1924, Costello 1954), and in the southern Rocky
Mountains between 2,440 m and 2,898 m
(Pearson 1931). The lower el- evational limits in
the central and southern Rocky Mountains vary
more from place to place than do the upper limits
(J.R. Jones,6 personal communica- tion). In
northern Colorado, absolute lower limits lie at
about 1,769 m on steep, north-facing slopes in the
lower foothills of the Front Range. Here, young
Douglas-fir forms dense stands on north-facing
slopes as low as 1,830 m. In some localities in
south- ern and central Arizona, Douglas-fir
descends to 1,830 m on north-facing slopes, and
may occur as low as 1,555 m in canyon bottoms.
But, in general, Douglas-fir is rarely found below
2,440 m in the Southwest.
In central Colorado, the species occurs as high
as
2,958 m on the rim of Coffee Pot Mesa; adjacent
vege- tation is sagebrush and fescue with aspen
groves. On the Frazer Experimental Forest in
northern Colorado (lat 40° N), a few overmature
Douglas-firs are scat- tered through a stand of
lodgepole pine with an understory of Engelmann
spruce and subalpine fir at an elevation of 2,267
m. In the White Mountains of east-central
Arizona, Douglas-fir is represented in Engelmann
spruce/corkbark fir stands up to 3,050 m
elevation. According to J.R. Jones (personal
commu- nication), the highest elevation at which
Douglas-fir occurs in the Southwest is at 3,264 m
on the crest of Mt. Graham in southeastern
Arizona.
In Mexico, Douglas-fir grows at altitudes of
2,000 to 3,200 m. Stands of Douglas-fir are
present largely in the Sierra Madre Occidental,
from Sonora and Chihuahua as far as Zacatecas,
in some mountain- ous area of Coahuila and
Nuevo León, in the highest part of the Sierra de
Pachuca, Hidalgo, and in a small
area in the center of the state of Puebla fir Subregion contained slightly more than half of
(Rzedowski and Huerta 1978). all standing Douglas-fir timber (Waddell et al.
1989). Oswald et al. (1986) listed the timber
Area occupied by Douglas-fir volume in the
The area occupied by Douglas-fir in the
United States was listed as 14.4 million ha in
1989—or 7.3% of the country’s 195.7 million
ha of non-reserved timberland (Waddell et al.
1989). The data on acre- age of forest types
dominated by Douglas-fir have changed over
time (Table 2.2). They show a decline of
3.6 million ha in the area of coastal Douglas-
fir from 1936 to 1977, and then an increase of
0.5 million ha by 1987. For interior Douglas-
fir, the data indicate a decrease of 1.8 million
ha from 1936 to 1952, fol- lowed by an
increase of 3.3 million ha in the next 35 years.
To what extent these changes represent actual
increases or decreases in acreage, and how
much of the changes reported reflect different
inventory procedures is open to question.
A breakdown of the area occupied by
Douglas- fir according to region and
productivity class (Table 2.3) indicates that
the acreage of Douglas-fir in the Pacific Coast
region exceeds that of Douglas-fir in the
Rocky Mountain region by 1.8 million ha.
The largest share of Douglas-fir in the Pacific
Coast region is concentrated in the Pacific
Northwest Douglas- fir Subregion, which
comprises western Oregon and western
Washington. Western Oregon con- tains
roughly two-thirds, and western Washington
one-third of the Douglas-fir acreage in the
Pacific Northwest Douglas-fir Subregion.
Moreover, that subregion has also the largest
share of acreage in the two highest
productivity classes.
The area occupied by Douglas-fir in
Canada, about 4.5 million ha is slightly less
than one-third of that in the United States.
According to the 1968 in- ventory data
(British Columbia Forest Service 1968),
Douglas-fir occupied 1.1 million ha in the
coastal region of British Columbia. The 1984
inventory data (British Columbia Ministry of
Forests 1984) show a decline of 15.2% in that
acreage to 900 million ha. Douglas-fir in
interior British Columbia occupied
3.6 million ha in 1968.
In 1987, the Pacific Northwest Douglas-
Chapter 2. Natural Range 23

Table 2.2 Area of Douglas-fir in the United States by variety and date of inventory.

Coastal variety a (thousand) Interior variety b (thousand) Total (thousand) Source


ha ha ha
1936 11,209 4,681 15,890 Mattoon 1936
1952 9,927 2,920 12,847 USDA 1958
1977 7,562 4,947 12,509 USDA 1982
1987 8,091 6,260 14,351 Waddell et al. 1989
a.Data from California, Oregon, and Washington.
b.Data from the Rocky Mountain region.

Table 2.3 Area occupied by Douglas-fir 1987 in the United States by region and productivity class (from Waddell et al. 1989).

Area (thousand ha)


Pacific Northwest Douglas-fir Pacific Southwest Ponderosa pine subregion Rocky Mountains
Productivity class subregion Totals
120+ 3,033 191 108 325 3,657
85-120 1,731 267 226 851 3,075
50-85 805 127 538 2,275 3,745
20-50 103 49 633 1,935 2,720
0-20 137 6 137 874 1,154
Totals 5,809 640 1,642 6,260 14,351

Table 2.4 Volume of standing Douglas-fir timber in the United States (Waddell et al. 1989).

Million
m3 ft3 %
Pacific Northwest 1,337 47,225 51.9
Pacific Southwest 360 12,701 13.9
Ponderosa pine subregion 242 8,570 9.4
Rocky Mountains 639 22,566 24.8
Total 2,578 91,062 100.0

Douglas-fir type of the Pacific Northwest Pacific Northwest has a truly maritime climate
Douglas- fir Subregion as 1.7 billion cubic char- acterized by mild wet winters and cool,
meters. That is, 328 million cubic meters (11.6 relatively dry summers, and long growing seasons.
billion cubic feet) more than given in the 1987 The climate becomes increasingly continental
inventory data by Waddell et al. (1989) (Table toward the east. The major environmental
2.4). According to Oswald et al. (1986), the gradients within the region are associated with
volume of coastal Douglas-fir in British Columbia distance from the ocean, latitude, and elevation.
amounted to 240 million cubic meters in 1984. In The Klamath area in southern Oregon and
1992, the combined volume for coastal and northern California are particularly hot and dry,
interior Douglas-fir was 687.6 million cubic whereas more northern areas on Vancouver Island,
meters for pure stands and 24.4 million cubic British Columbia, and the Olympic Peninsula of
meters in mixed stands (Macklin and Manning Washington include temperate rainforests with up
1992). to 4,445 mm of rain a year. Most of the
precipitation occurs in winter as rain although
Climate snow is prevalent at higher elevations particularly
Douglas-fir grows under a wide variety of in the Cascade Range and the Sierra Nevada.
climatic conditions (Table 2.5). The coastal
region of the
24 Douglas-fir: The Genus Pseudotsuga

Table 2.5 Climatic data for five subdivisions of the range of Douglas-fir.

Pacific Northwest Rocky Mountains


Climatic data Coastal Mountainous Northern Central Southern
Mean temperatures (°C)
July 20–27 22–30 14–20 14–21 7–11
January −2.5 to 2.5 −9.0 to −2.5 −7.0 to −2.5 −9.0 to −6.0 0 to 2.0
Frost-free period (d) 195–260 80–180 60–120 65–130 50–110
Precipitation (cm)
Mean annual 76–300 60–300 56–102 36–61 41–76
Snowfall 0–60 10–300 41–584 50–460 180–300

In the northern Rocky Mountains, Douglas-fir the order Inceptisols have formed on the younger
grows in a climate with a marked maritime influ- glacial landscapes. Alfisols (Gray Brown Podzolic
ence, except for a dry period in midsummer. In soils) and Ultisols (Red-Yellow Podzolic soils)
the central Rocky Mountains, Douglas-fir are found on older surfaces. Because cooler and
experiences a continental climate. Winters are wetter climates at higher elevations promote
long and severe; summers are hot, and, in some podsolization, Spodosols (Podzols) are found.
parts of the region, very dry. Annual precipitation, Soils farther inland within the range of the coastal
which is higher on the western sides of the variety are derived from a wide variety of parent
mountains, is mainly snow. Rainfall patterns for materials. They include metamporphosed
the southern Rocky Mountains generally show sedimentary material in the north- ern Cascades
low winter precipitation east of the Continental and igneous rocks and formations of volcanic
Divide, but high precipitation during the growing origin in the central and southern Cascades. Depth
season. West of the Continental Divide, the of soils ranges from very shallow, on steep slopes
rainfall is more evenly divided between winter and ridgetops, to deep, where deposits of
and summer. Frost may occur in any month in the volcanic origin and residual and colluvial
northern part of the range. The length of the frost- materials are found. Texture varies from gravelly
free period, however, varies within the central and sand to clays. Organic matter content ranges from
southern Rocky Mountain regions, even at the moderate in the Cascade Range to high in parts of
same elevations. the Coast Range and Olympic Peninsula. Total N
Soils content varies con- siderably but is usually low in
soils of glacial origin. Great soils groups
The coastal variety of Douglas-fir reaches its best characteristic of the range of Douglas-fir include
growth on well aerated soils with a pH range from Haplohumults (reddish Brown Lateritics) of the
5 to 5.5. It will not thrive on poorly drained or order Ultisols, Dystrochrepts Brown Lateritics),
com- pacted soils. Soils in the coastal belt of Haplumbrepts (Sols Bruns Acides) of the order
northern California, Oregon and Washington Inceptisols, Haplorthodes (Western Brown
originated chiefly from marine sandstones with FVorest Soils) of the order Spodosols,
scattered igneous intrustions. These rocks have Xerumbrepts (Brown Podzolic soils), and
weathered deeply to fine-textured, well-drained Bitrandepts (Regosols)
soils under the mild, humid climate of the coast. (Heilman et al. 1979).
Surface soils are in general moderately acid, high Soils influenced by volcanic ash (Andepts)
in organic mat- ter and total nitrogen, and low in occur throughout the Cascades of Washington and
base saturation. Soils in the Puget Sound area and Oregon. An extensive area of such soils is on the
in southwestern British Columbia are almost east side of the Cascade Range in southern
entirely of glacial origin (Tarrant 1956). Oregon. Along the Pacific coast and in
Podzolization is the dominant soil- forming southwestern Oregon, except at high altitudes,
process but is not intensive in the mild, moist older landscapes and a warmer climate give rise to
climate. Inceptisols (Sols Bruns Acides) of Ultisols (Reddish-Brown Lateritic soils). East of the
Cascade Range where a more severe and arid
climate prevails, Alfisols (Brown Forest
Chapter 2. Natural Range 25
mainly eliminated the extensive original old-growth
soils and Gray Wooded soils) are common except forest, clearcutting combined with slash
for some higher elevations where Spodosols have
developed. Farther east, Alfisols grade into
Mollisols (Chernozems, Rendzinas) which cover
the extensive arid areas of interior Washington
and Oregon.
Soils within the range of the interior variety of
Douglas-fir also originated from a considerable
array of parent materials. In south-central British
Columbia, eastern Washington, and northern
Idaho, soils vary from basaltic talus, to deep loess
with volcanic ash, to thin residual soils over
granitic or sedimentary rocks. They are mostly
Vitrandepths and Xerochrepts. Parent materials in
Montana and Wyoming consist of both igneous
and sedimentary rocks, and locally of glacial
moraines. Soils derived from noncalcareous
substrates are variable in texture but are
consistently gravelly and acidic. A signifi- cant
portion of the sedimentary rock is limestone,
which gives rise to neutral or alkaline soils rang-
ing in texture from gravelly loams to gravelly
silts. Limestones often weather into soils that are
exces- sively well drained. Soils are Cryoboralfs
of the order Alfisols, and Cryandepts and
Cryochrepts of the order Inceptisols. Soils in the
central and southern Rocky Mountains are very
complex. They developed from glacial deposits,
crystalline granitic rocks, con- glomerates,
sandstones, and, in the Southwest, lime- stones.
These soils are Alfisols (Gray Wooded soils),
Mollisols (Brwon Forest soils), Spodosols (Brown
Podzolic soils, Podzols), and Entisols (Alexander
1974, Pfister et al. 1977).
Associated forest cover
Douglas-fir grows together with other conifers
and hardwoods throughout its natural range. The
kinds of species mixture are listed by the Society
of American Foresters as forest cover types (Eyre
1980). Coastal Douglas-fir is the dominant
component of type 229 (Pacific Douglas-fir). This
type is restricted in the United States to areas west
of the Cascade Range in Washington and Oregon,
and to a more limited area in northwestern
California. Periodic recurrence of catastrophic
wildfires created vast al- most pure stands of
Douglas-fir throughout its range north of the
Umpqua River in Oregon. Although logging has
Douglas-fir is a component of Sierra Nevada
burning has helped maintain Douglas-fir as mixed conifers (Type 243). Five conifers define
the major component in second-growth stands Type 243: California white fir, Pacific ponderosa
(Hermann and Lavender 1990). Toward the pine, sugar pine, incense cedar, coastal Douglas-
fog belt of the Pacific coast, Douglas-fir gives fir, and black oak. Douglas-fir dominates the mix
way to types 230 (western hemlock), 223 in the north, but is absent south of the Merced
(Sitka spruce), 228 (western red cedar) but River (Laacke and Fisk 1983).
remains a common component of these types.
Douglas-fir is usually an early seral
component of forests. Large continuous
stands are succeeded by more shade-tolerant
species, especially western hemlock, that
regenerate and grow better below the canopy
of mature Douglas-fir, unless natural catas-
trophes, such as wildfire and windthrow,
intervene.
South of type 229 (Pacific Douglas-fir),
there is a transition to types 234 (Douglas-fir-
tanoak-Pacific madrone) and 244 (Pacific
ponderosa pine-Douglas- fir, the mixed
conifers and hardwoods of southwest- ern
Oregon. Mixed conifer forests occupy nearly
half of the area of southwest Oregon. They
extend from the Calapooya Mountains
(143°30’ N) south into northwestern
California, and from the western slopes of the
Cascade Range to the Pacific Ocean (Minore
and Kingsley 1983). The mixed-conifer
forests of southwestern Oregon vary in
composition, but two or more of the following
species are always present: Douglas-fir, incense
cedar (Calocedrus decur- rens), sugar pine
(Pinus lambertiana), white fir (Abies concolor),
and ponderosa pine (Pinus ponderosa).
Hardwoods often associated with mixed
conifers include giant chinkapin (Chrysolepis
chrysophylla), California black oak (Quercus
kelloggii), Oregon white oak (Quercus
garryana), canyon live oak (Quercus
chrysolepis), and bigleaf maple (Acer
macrophyllum). In the past, wildfires tended to
perpetuate seral species, such as Douglas-fir,
ponderosa pine, incense cedar, and Pacific
madrone (Arbutus menziesii) throughout
southwestern Oregon. Under modern fire
prevention practices, however, natural stands
slowly convert to more shade-tolerant species,
such as white fir, Port Orford cedar
(Chamaecyparis lawsoniana), and tanoak
(Lithocarpus densiflorus) (McDonald et al.
1983).
26 Douglas-fir: The Genus Pseudotsuga
Douglas-fir is the principal component of forest from the Figueroa and San Emigdio Mountains
cover type 210 (interior Douglas-fir), apart from southward to its type locality in Banner
grand fir (Abies grandis), ponderosa pine, lodgepole
pine (Pinus contorta), and western larch (Larix
occidentalis), of for- ests in most of eastern
Oregon, eastern Washington, and adjacent British
Columbia. Douglas-fir is the cli- max species in
these forests north of the Entiat River in
Washington. Both Douglas-fir and grand fir form
climax forests on the east side of the Cascade
Range in Washington south of the Entiat River.
Grand fir is the usual climax species in Oregon
(Franklin and Dyrness 1973). Stand structure and
species compo- sition of the stands formed by
Douglas-fir and its associates in eastern
Washington and eastern Oregon are extremely
variable, depending upon site, past management
practices, and fire history.
Douglas-fir grows in extensive pure stands,
either
even- or uneven-aged, or in mixtures with pon-
derosa pine, western larch, grand fir and
lodgepole pine, in southern Idaho, northern Utah,
and western Montana. These forests are
represented by forest cover types 210 (interior
Douglas-fir, 237 (interior ponderosa pine) 212
(western larch) and 213 (grand fir). Wherever
Douglas-fir grows in mixture with other species,
the proportion may vary greatly, de- pending upon
elevation, aspect, soil, and especially fire history.
That is particularly true of the mixed conifer
stands in the central and southern Rocky
Mountains (Alexander 1974). While some stands
may consist of only two species, others may be
composed of as many as seven additional
associates, along with Douglas-fir: ponderosa
pine, white fir, Engelmann spruce (Picea
engelmannii), southwestern white pine (Pinus
strobiformis), corkbark fir (Abies lasiocarpa), blue
spruce (Picea pungens), and quaking aspen
(Populus tremuloides). Stands are often multistoried,
with Douglas-fir and interior ponderosa pine in
the overstory (Muldavin et al. 1996, Hoffman and
Alexander 1983, Moir and Ludwig 1979).
Pseudotsuga macrocarpa
The distribution of bigcone Douglas-fir,
Pseudotsuga macrocarpa (Vasey) Mayer, is
restricted to south- ern California. Its
discontinuous range forms an arc extending
Canyon at the southern end of the Volcan environment of northern Baja California appears
Mountains in San Diego County (Figure 2.6). to be unsuitable for members of the evergreen
Bigcone Douglas- fir occurs in all mountain mixed forest, such as P. macrocarpa, that have
ranges, except for the Hot Springs, Santa mesic temperature requirements.
Rosa, and northern San Jacinto Mountains,
which are relatively arid, owing partly to rain
shadows from the Cuyumaca, Palomar, and
Santa Ana Mountains to the west (Minnich
1982). According to McDonald (1990), the
northern limits of the range of the species are
near Mt. Pinos in Kern County, and the
headwaters of Labrea Creek in Santa Barbara
County. As westernmost limits, he indicates
Mission Canyon in the Santa Ynez
Mountains, and Zaca Peak in the San Rafael
Mountains.
Some older publications (Bergen 1904,
Sudworth 1908, Standley 1920-26, Davidson
and Moxley 1923, Bowers 1942, Dallimore
and Jackson 1948) contain references to the
presence of Pseudotsuga macrocarpa in Baja
California, Mexico. Both Munns (1938) and
Gause (1966) mapped bigcone Douglas-fir in
the Sierra de Juárez and the Sierra de San
Pedro Mártir of Baja California south to lat
31° N—that is, about two degrees farther
south than Banner Canyon.
Doubts about a Mexican distribution of the
spe- cies have long existed. Martinez (1949)
emphasized that he had not encountered
bigcone Douglas-fir in Baja California or any
other region of Mexico. The question of
whether P. macrocarpa is present in Baja
California appears to have been conclusively
answered by Minnich (1982). His search for
the tree in Baja California, both on the ground
and by means of aerial photographs, was
unsuccessful. Although he conceded that
absolutely disproving the occur- rence of
bigcone Douglas-fir in such inaccessible
country is impossible, he provided additional
argu- ments for the absence of the species in
Mexico. One is the likelihood that Sudworth’s
(1908) account of bigcone Douglas-fir in Baja
California is based on either a
misinterpretation of a geographic name or
reliance on an ambiguous report by North
(1907) on conifers in the Sierra de San Pedro
Mártir; this set off a chain reaction of
erroneous references. The second and even
more compelling argument is that the physical
Chapter 2. Natural Range 27

119° 117° 115°

Figueroa Mountain San Emigdio Mountains

San Rafael Mountains

Santa Barbara
Transverse Ranges

N Los Angeles

California

Volcan Mountains

Santa Isabel
Banner Canyon

33°
San Diego

Tijuana Mexicali
Baja California

Ensenada
Isolated populations
Numerous fragmented populations mostly of
Continuous populations extending at least 5 km
050 Km small size

San Felipe

33°

Figure 2.6 Range of Pseudotsuga macrocarpa (from Minnich 1982).

The altitudinal distribution ranges from 275 m


on cool, moist north slopes of canyon bottoms to 1,830 m, bigcone Douglas-fir grows in open stands
2,400 m on warm south-facing plateaus in mixture with Pinus ponderosa, Pinus jeffreyi,
(McDonald 1990). In the Transverse Ranges, Pinus lambertiana, and Calocedrus decurrens
larger stands are found from 915 m to 1,650 m on (Gause 1966).
southwest through north slopes, mainly in the Pseudotsuga japonica
upper canyons. Vigorous stands occur in the
The distribution of Japanese Douglas-fir,
Coastal and Peninsular Ranges in westerly
Pseudotsuga japonica (Shinas.) Beissn., is limited to
canyons from 730 m to 1,525 m, apparently
Japan. The spe- cies occurs on the Kii Peninsula of
because of the year-round influx of marine air.
southern Honshu and on Shikoku (Figure 2.7).
Above
Hayata (1915) pointed
28 Douglas-fir: The Genus Pseudotsuga
132° 134° 136°
36°

Abundant Less abundant

JAPAN

Kobe Nara
Okayama Osaka
Hiroshima
Takamatsu Kamikitayama
Tokushima Wakayama
34°
34°
Matsuyama
Kii Peninsula
Kochi

Shikoku Island
0 50 100 km

0 50 100 mi

132° 134° 136°


Figure 2.7 Range of Pseudotsuga japonica Beissn. (from Hayashi 1952).

out that he and other botanists had erroneously because the litera-
regarded P. japonica and P. wilsoniana as
identical, which explains why he indicated earlier
(1905) the presence of P. japonica on Taiwan.
The Japanese Douglas-fir is a rare tree whose
range is highly discontinuous (Hayashi 1952).
The northern limit of the species is on Mt.
Kunimi at lat 32°22’ N, longitude 136°10’E.
The range extends southwestward through Mie,
Nara, and Wakayama Prefectures on the Kii
Peninsula. The tree occupies only a small area
in the southeastern part of Shikoku Island. Its
western limit is about longitude 134°05’E; the
southernmost occurrence is on Mt. Senbon in
Kochi Prefecture at lat 33°26’N. Vertical
distribution extends from 400 m to 1,000 m on
Shikoku Island and to 1,100 m on the Kii
Peninsula. Most of the trees, however, grow at
elevations between 500 m and 900 m (Ohwi et
al. 1965).

The Chinese Douglas-firs


Descriptions of the ranges of the Chinese
Douglas- firs may not be entirely accurate
ture contains conflicting statements about
taxonomic status and distribution.

Pseudotsuga wilsoniana
The Formosan Douglas-fir, Pseudotsuga
wilsoniana (Hayata), is limited to Taiwan and
a few locations in China (Figure 2.8). In
Taiwan, the species extends north-south from
about lat 24°41’ N to 23°03’ N, and east-west
from about long 122°24’ E to 121°24’ E (Liu
1966). The tree is rare (Li 1950) and is
restricted to a belt between 800 and 1500 m in
steep, mountainous terrain (Lin et al. 1953).
Li (1975) cites 2,500 m as the upper
altitudinal limit of the species. Liu (1966)
listed
P. wilsoniana as endemic in Formosa.
According to Wang (1961), however, a few
relicts of the species exist in the southern part
of Fujian province in China. Most of the
forest in this province has been cleared, and
only remnants of the former forest still exist.
At higher elevations, these remnants
supposedly contain Formosan Douglas-fir.
The island of Taiwan was separated from
the Chinese mainland during the Pleistocene
200,000 to
Chapter 2. Natural Range 29

400,000 years ago (Liu 1966). The proximity of Yangtze region. He described the distribution
Fujian and Taiwan suggests that the Formosan (Figure 2.8) as “widely scattered in west Hupeh
Douglas-fir may once have formed a continuous [Hubei], northwest and south Hunan, northeast
range in the two regions. Kweichow [Guizhou] and southeast Szechuan
[Sichuan] at an elevation of 800 to 1,200 m. In
Pseudotsuga sinensis
southwest Szechuan, central and northeast
Wang (1961) wrote that the Chinese Douglas-fir, Yunnan, the tree occurs at 1,500 to 2,800 m
Pseudotsuga sinensis (Dode), extends over a altitude.”
distance of nearly 2,000 km along the Yangtze
Valley from the Pacific coast to western Sichuan. Pseudotsuga gaussenii
But according to Chengde (1981), the species is Eastern Chinese yellow fir, recognized by Chinese
limited to the upper botanists as a species separate from Pseudotsuga

RUSSIA

Heilongjiang
KAZAKHSTAN

MONGOLIA Jilin
KYRGIZSTAN

Liaoning NORTH KOREA

Xinjiang
Inner Mongolia (Nei Mongol) Beijing
Gansu Tianjin
Hebei
SOUTH KOREA
PAKISTAN
Ningxia Shanxi
Shandong
CHINA
Qinghai
Henan Jiangsu
Tibet (Xizang) Shaanxi
Shanghai
Anhui
Hubei Chongqing
Sichuan
NEPAL BHUTAN Zhejiang
Hunan Jiangxi
Fujian
Guizhou

INDIA Yunnan Guangxi Guangdong TAIWAN


Hong Kong Macau
BANGLADESH

MYANMAR (BURMA) VIETNAM


LAOS
Hainan Locations of Pseudotsuga
P. forrestii
P. gaussenii
THAILAND P. brevifolia
P. sinensis
CAMBODIA
P. wilsoniana

Figure 2.8 Range of Pseudotsuga in China (after Chengde 1981).


30 Douglas-fir: The Genus Pseudotsuga
sinensis, grows in the lower Yangtze region fir were comments on herbarium sheets of speci-
(Figure 2.8). Wang (1961), who still called the mens collected in 1914 by Forrest (Craib 1919) in
Douglas-fir of that region P. sinensis, listed it as the Mekong-Salween watershed of northwestern
occurring in northern Zhejiang, southern Anhui, Yunnan at lat 27°40’ N, and in 1922 by Maire
and northern Jiangxi. Chengde (1981) described (Wilson 1926) in southeastern Tibet at lat 28°25’
the distribution of P. gaussenii as “only scattered N. Chengde (1981), apparently following Cheng
in south Anhwei [Anhui], and west and south and Fu (1978), gave the range of Pseudotsuga
Chekiang [Zhejiang] at 600 to 1,500 m altitude.” forrestii Craib as in- cluding northwest Yunnan,
Cheng and Fu (1978) re- ported that P. gaussenii southeast Tibet, and southwest Sichuan, at
occurred also in the province of Guangdong. elevations of 2,400 to 3,300 m.

Pseudotsuga forrestii Pseudotsuga brevifolia


Until the publication of the keys to Chinese Shortleaf yellow fir has a limited distribution in
Douglas- firs by Cheng and Fu (1978), the only Longzhou and Jingxi counties of southwest
available re- cords on the distribution of the Guangxi at altitudes of about 1,250 m (Chengde
Mekong yellow 1981) on south slopes or near peaks (Cheng and
Fu 1978).
3. Areas of Introduction
Richard K. Hermann

P
seudotsuga menziesii is distributed more
widely outside its natural range than any Northern Hemisphere
other species of American forest tree, with The principal region of introduction of
the exception of Douglas-fir to the northern hemisphere is
Pinus radiata. Its successful introduction beyond Europe. By compari- son, the extent of
its natural habitat into many parts of the introduction to regions in North America
temperate regions of the northern and southern outside the natural range of the species is minor.
hemisphere is all the more remarkable because
of the ignorance of, or disregard for, the Western North America
importance of provenance variation until lately. Alaska
The introduction of Douglas-fir went through
The current range of P. menziesii var. menziesii
various phases. Initially the species was intro-
does not extend into Alaska although fossil evi-
duced through individual tree plantings in
dence indicates that Douglas-fir grew there in the
Europe and elsewhere around the world.
Miocene (Wolfe 1969) and Pleistocene (Hopkins
Successes were mostly dependent on the seed
and Benninghoff 1953). A few small plantations
sources and setbacks were due to the occurrence
with Pseudotsuga menziesii var. menziesii were
of diseases, especially Rhabdocline pseudotsugae
established, one as early as 1927, in southeast
and Phaocryptopus gaeuman- nii. Over time, seed
Alaska (Harris 1971) from 208 km to 352 km north
source problems and diseases were overcome
of the coastal variety’s northern natural limit.
through genetic selection and silvi- cultural
Although experience with Douglas-fir in Alaska is
practices that allowed for wider establish- ment,
limited, Harris stated that, “under present climatic
including monoculture stands. Social issues
conditions the species is capable of germinating,
influencing such phases included trade barriers,
becoming established, mak- ing excellent growth,
e.g., during war time, but more influential were
and producing viable seed far north of its present
the political discussions about the introduction
northern coastal natural limit.”
of non- native species. The following chapter
describes this development until the mid-1990s. Hawaii
At present, Douglas-fir is an accepted and inte-
The compilation of “forest plantings in Hawaii by
gral part of forest management in many countries
genera, 1908-1960” (Nelson 1965) lists the num-
because of its economic importance and its repu-
bers of Douglas-fir planted as 1,835 trees, but
tation as a species that may be better able to deal
does not provide exact dates and locations of
with climate change, especially with its drought
plantings. Included in that number is probably a
resistance. However, the proportion of Douglas-
small planta- tion established by L.W. Bryan, a
fir is often limited because of concerns about its
former Territorial Deputy Forester for Hawaii. He
eco- logical impacts; for example Forest
planted 50 seedlings of coastal Douglas-fir in
Stewardship Certification standards in Germany
1934 near the 1,829 m level on the northeast slope
limit non-native species to 20% stocking in
of majestic Mauna Kea as a
management units.
31
32 Douglas-fir: The Genus Pseudotsuga
monument to David Douglas. At this spot in 1834, To explore the feasibility of growing Douglas-
Douglas met a violent death in a wild bullock pit. fir as Christmas trees in Pennsylvania, a series of
Of the 50 trees planted in 1934, some were plantings were made at eight locations in the state
uprooted by wild pigs and others smothered by beginning in 1952 (Bramble and Byrnes 1952). Of
vines. In 1976, the largest of the surviving trees the 19 coastal and interior provenances used in the
had a height of 37 m and a dbh of 76 cm. Coastal study, those from the Pacific Northwest suffered
Douglas-fir in that grove, situated at about lat. severe winter injury. After six years, 55% of the
10°30’N, represents trees of the variety menziesii coastal Douglas-firs had died (Byrnes et al. 1958).
growing closest to the equa- tor (Nelson 1976). In In an effort to find Douglas-fir provenances
1984, on the 50th anniversary of the original suit- able for Christmas tree plantations,
planting of Douglas-fir at that site, a second shelterbelts, and ornamental plantings in the
planting of coastal Douglas-fir took place nearby Midwest, a major study was initiated in 1961
(Anonymous 1985). (Wright et al. 1971). Trees for the study were
grown in a nursery near East Lansing, Michigan,
Central and Eastern North America from seed collected in 128 locations throughout
the natural range of the species in the United
Attempts to grow the coastal variety east of the
States and Canada. Subsequently, seed- lings were
Rocky Mountains have mostly failed. Jäger and
distributed to participants in the study to establish
Beissner (1884) mentioned a communication from
test plantations in Nebraska, Michigan, and
A. Strauch, Superintendent of Spring Grove
Pennsylvania. Seedlings of P. menziesii var. men-
Cemetery in Cincinnati, Ohio, to the effect that
ziesii had already suffered extreme winter injury
coastal Douglas- fir would do extremely poorly
in the East Lansing nursery, and many of them
in the central and eastern United States. Sargent
died in the nursery. Gerhold (1966), who tested
(1898) wrote: “Early attempts to introduce it into
67 of the provenances from the Wright collection
the eastern United States by means of plants
in a nursery near Potters Mills, Pennsylvania, also
obtained in England and raised from seeds
reported severe damage by winter cold to
gathered in Oregon, or from trees which had
seedlings from west coast seed sources. Of the
grown in Europe, were generally unsuccessful,
trees of 14 west coast provenances from the
the young plants soon succumbing to the heat
Wright collection grown in a Nebraska test
and dryness of the eastern summers or to the
plantation, all from 12 of the 14 provenances died
cold of eastern winters.”
in the first three years after outplanting. Read and
C.A. Schenck established a small plantation of
Sprackling (1976) concluded that trees of the
Douglas-fir with 4-year-old plants in 1896 on the
coastal variety of Douglas-fir cannot survive
Vanderbilt estate near Biltmore, North Carolina.
Nebraska winters, and hence, should not be
Trees were 5.5 to 7.3 m high with diameters
planted there.
breast high of 8.9 to 12.7 cm at age 32 from seed. Pseudotsuga menziesii var. glauca has been grown
The plan- tation did not thrive, however, because as an ornamental tree for more than 100 years in
of infection with Polyporus schweinitzii the eastern United States. Sargent (1898) wrote:
(Hedgcock et al. 1925).
In 1862 Dr. C. C. Parry found the Douglas Spruce on
None of the Douglas-firs in a plantation estab-
the outer ranges of the Rocky Mountains of Colorado,
lished in 1919 in Mahoning County, Ohio, and the following year sent seeds to the Botanic
survived (Aughanbaugh 1960), but Douglas-fir Garden of Harvard College. The plants raised from
planted 1928 at Cloquet in northeastern these seeds have proved perfectly hardy and have
Minnesota about 30 km (18 mi) west of Lake grown rapidly and vigorously in the neighborhood of
Superior had 90% survival at age 20, and 83% at Boston, and now give promise of surpassing all other
exotic conifers in permanent beauty and usefulness;
age 41 (Alm et al. 1972). Grigsby (1969) reported
and in recent years the Douglas Spruce, raised from
on performance of non-native spe- cies under
seeds gathered at high altitudes in Colorado, has been
planting conditions in southern Arkansas and planted in considerable numbers in the northern states.
northern Louisiana. Among species that failed
completely was P. menziesii of California origin.
Chapter 3. Areas of Introduction 33

Results from the test plantations established Europe


with trees of the interior variety from the Wright
Weck (1949) wrote, “Currently more pure stands
prov- enance collection have pointed to various
of Douglas-fir, introduced to Europe around 1700,
degrees of adaptability between provenances to
exist than any other introduced species.” 1 Volk
growing conditions outside their natural range. In
(1969) claimed that Douglas-fir was brought to
a 12-year- old plantation at Kellogg, Michigan,
Germany at the end of the 18th century by Baron
provenances from Arizona, New Mexico, and
von Wangenheim, a Hessian officer sent to
Colorado proved to be highly susceptible to late
America to fight for the British in the
spring frost injury, whereas provenances from the
Revolutionary War. Both these accounts of the
northern Rocky Mountains in Montana and Idaho
introduction of Douglas-fir to Europe are
were not sus- ceptible (Steiner and Wright 1975).
erroneous.
Van Haverbeke (1987), summarizing the 20-year-
The event that marked the introduction of P.
performance results of interior provenances at the
menziesii to Europe was the arrival in early 1827
Plattsmouth plantation in eastern Nebraska, found
of the Douglas-fir seed shipped to Great Britain
that they were in close accord with those reported
by David Douglas. In the nearly 200 years since
at age 11 by Reed and Sprackling (1976). More of
arrival of the first shipment of seed, Douglas-fir
the trees from southern Rocky Mountain
has become more widely distributed in Europe
provenances survived and were taller after 20
than any other North American conifer (Figure
years than those from central and northern Rocky
3.1).
Mountain provenances. The good showing of the
Both varieties of P. menziesii have been
trees of southern Rocky Mountain origin is
planted in Europe, but the coastal variety turned
remarkable because they suffered heav- ily from
out to be far better suited to cultivation in most
repeated dieback of terminal shoots in successive
European forest regions than has the interior
winters caused by cold injury. Superior
variety. Based on an analysis of climate and
performance of New Mexico provenances over
physiography, Schwarz in 1933 concluded that the
those from the central and northern Rocky
potential range for cultivating Douglas-fir in
Mountains had also been noted in earlier trials in
Europe would include the southernmost parts of
New Hampshire (Baldwin and Murphy 1956) and
Norway and Sweden, Denmark, Germany,
Iowa (Erdmann 1969). Douglas-fir in a mixed
western Poland, parts of north- ern Austria,
conifer plantation established 1960 in
northern Switzerland, Belgium, Holland, Great
Newfoundland (Singh 1970) and in a small
Britain, Ireland, northern Spain, Portugal, and
plantation made in 1941 on Prince Edward Island
France exclusive of its Mediterranean region. His
(Peterson 1964) are the easternmost introductions
assessment of conditions for growth of coastal
in North America on record. Whether these trees
Douglas-fir in various parts of Europe, although
belonged to the coastal or interior variety of
founded mainly on theoretical considerations, was
Douglas-fir was not indicated.
Experience over the span of a century clearly proven to be largely correct by more than a
shows that the coastal variety of Douglas-fir is un- century of experience with the cultivation of
suitable for planting in North America east of the Douglas-fir.
variety’s natural range because they lack frost Great Britain
hardi- ness. Trees of the interior variety can better
That the introduction of Douglas-fir to Europe be-
adapt to climatic conditions in central and eastern
gan in the British Isles is probably more than just
North America, although their potential for
a historical accident. As noted by Macdonald
survival and growth varies considerably between
(1957), exotic trees play a more important role in
provenanc- es. A serious threat, especially in the
Britain than in other European countries because
northeastern United States, to cultivation of the
of the poverty
variety glauca is its great susceptibility to
infection by the fungus Phaeocryptopus gaeumannii,
which causes needle cast. 1. Translated from the original: “Fuer die bereits um 1700 nach
Europa gebrachte Douglasie liegen gegenwärtig mehr
Einzelanbauflächen vor als für jede andere eingeführte
Holzart” (Weck 1949, p. 20).
34 Douglas-fir: The Genus Pseudotsuga

20° 0° 20° 40°

0 300 km

0 300 mi

ICELAND

FINLAND
70 (1995)

NORWAY RUSSIA

60° ESTONIA

SWEDEN
100 (1995)
LATVIA

LITHUANIA
BELARUS
3,000 (1990) RUSSIA
DENMARK

IRELAND

47,000 (1985) THE


POLAND
8,000 (1993) UNITED
NETHERLANDS GERMANY UKRAINE
18,400 (1996) 100,000 (1995) 1000 1,550 (1985)
KINGDOM
(1993)
BELGIUM
CZECH

18,000 (1994) REPUBLIC


3500 (1991) SLOVAKIA
1200 (1993)
LUXEMBOURG HUNGARY ROMANIA
AUSTRIA 353 13,000 (1993)
1,000 (1995) (1990)
SWITZERLAND
SLOVENIA CROATIA
BOSNIA/ SERBIA BULGARIA
HERZEGOVINA 12,664 (1985)
FRANCE
MONTENEGRO
333,000 (1993) ITALY
MACEDONIA

10,000 (1985) ALBANIA

SPAIN
GREECE
40° 7,000 (1989) 30,000
PORTUGAL (1993) 100 (1985)

i
S

Figure 3.1 Areas ocupied by Douglas-fir in Europe, in hectares (modified from De Champs 1997b). Some smaller stands or
plantings may not be shown on this map, but may b ediscussed in the text or may appear on individual country maps.
Chapter 3. Areas of Introduction 35

of its native arboreal flora. That is particularly the Santa Cruz Mountains of California, and from
true of conifers whose sole indigenous collections for the Oregon Association made by
representatives are Scots pine (Pinus sylvestris), William Lobb and John Jeffrey in 1852/53
English yew (Taxus baccata), and juniper (Edwards 1957). Lobb who had come to
(Juniperus communis). The ac- quisition of California in the sum- mer of 1849 made an
colonies in different parts of the world favored excursion three years later to the Columbia River
acquaintance with foreign plants and their and Oregon, where he succeeded in obtaining
subsequent import. By the time Douglas-fir came seeds of Douglas-fir (Dallimore 1932). Some of
to Britain, people of that country had considerable the seed obtained by Lobb may have come from
experience with exotics, and displayed none of the locations not far from those of the Douglas
resistance to their introduction sometimes encoun- collections. The origin of the Douglas-fir seed
tered on the continent. sent by Jeffrey is unknown. Most likely, the seed
The first Douglas-firs grown in Britain arose came from stands in California because he is not
from seed sent by David Douglas in the fall of known to have made collections farther to the
1826. The exact source of that seed is unknown, north.
although a common assumption is that the seed Trees from Douglas’ original seed grew so
was from collec- tions Douglas made near his well that they excited considerable interest, and
base at Fort Vancouver on the Washington side when they began to bear cones, seed was collected
of the Columbia. Booth (1890, p. 47) reasoned as from them. Loudon (1838, p. 2321) wrote: “The
follows: “Because he ar- rived on August 31 at tree bore cones for the first time in England, at
Fort Vancouver, and, accord- ing to his diary, Dropmore (Buckinghamshire) in 1835, when the
sent already on September 1 his boxes on board plant there already mentioned produced one cone.
of the ship that left that same day for England; This year (1837) it has above a dozen; so that in
his diary entries from 2 to 9 September ‘laid in all probabil- ity, there soon will be an abundance
specimens of Pinus taxifolia with fine cones’ lead of seeds of this species, from which extensive
to the conclusion that he could have collected plantations may be raised, and the value of the
these only in southern Washington on the border species as a timber tree proved.”
to Oregon after he had sent off his collection from The most famous, however, are two trees at
northern Washington on the border to Oregon. Lynedoch, Scone Estate in Perthshire, planted in
Until proof to the contrary, I state on the basis of 1834; one of them was a particularly heavy cone
Douglas’ diaries that the trees from the year 1826 bear- er. The first cones were harvested from
described by me earlier must have originated these trees in 1844. In the next 30 years, the two
from the seed collected near Vancouver at the trees yielded about 200,000 cones from which
border to Oregon.” The Royal Horticultural about 4 million plants were raised (Booth 1877, p.
Society, sponsor of Douglas’ expedition, 63). They are known to be the source of many
distributed the seed from his 1826 shipment stands in Perthshire and some in Argyll (Lines
among its members. Trees raised from that seed 1987), and are still in existence.3 The seed lots
were planted in the parks of numerous estates. sent by Hartweg and Lobb produced many
Hutchison (1873) lists sites of early plantings of specimen trees but did not produce such vigorous
Douglas-fir and remarks in that context “the tree progeny as the seed of Douglas (Matthews 1953).
at Raith, near Kirkcaldy, in Fife, planted by Booth (1890, p. 43) mentions that he saw about 50
Douglas himself is now a splendid specimen.” splendid Douglas-fir at Murthly Castle raised
Details on locations of early plantings are also from the seed collected by Stewart at the end of
provided by the 1840s in California. In 1871, seed worth 75
A. Murray (1884), Elwes and Henry (1909), and pound sterling was collected from these trees.
Anderson (1967). Many of these trees still existed Booth emphasized
in the 1950s (Edwards 1957, Streets 1962), some in
1980 (Zander 1980), and others still in 1993.2
Other early introductions were from seed collect- ed in the late 1840s by Sir William Douglas
Stewart in California (Booth 1890), in 1846/47 by 2. Letter from Dr. D.C. Malcolm, University of Edinburgh, dated
Hartweg in 7 December 1993.
3. Alan Fletcher, British Forestry Commission,
personal communication.
36 Douglas-fir: The Genus Pseudotsuga
that plantations established from that seed were vol- ume had increased to 339 cubic meter/ha (Crozier
better than those from imported seed. 1908). Subsequent growth of the plantation appears
British land owners became interested in
Douglas- fir as a forest tree relatively early.
Douglas-fir, like other exotics, had been tried first
on fertile, sheltered sites where it had displayed
rapid growth (Anderson 1967). The vigorous
growth of the species in parks and arboreta
prompted some estate owners to plant it under
forest conditions in small plots. The remark- able
performance of Douglas-fir on such trial plots led
to establishing a 2.4 ha plantation in 1858 on the
estate of the Duke of Montrose in Buchanan,
Stirlingshire (Anderson 1967). Another, and
perhaps the best known of the early Douglas-fir
plantations, is the one at Taymount on the former
Mansfield estate in Perthshire. William
McCorquodale (1880), who worked as wood
surveyor on the Mansfield estate, gives the
following account of the plantation’s estab-
lishment: “Again in 1860, 8 acres were enclosed
with rabbit-proof wire netting on the estate of
Taymount, alongside the Highland Railway, and
planted with Douglas firs at 9 feet apart for the
permanent crop, and the intermediate spaces were
filled up with larch as nurses. This plantation is
now 20 years of age, and the nurses are all thinned
out. It now stands a pure Douglas fir plantation, in
prime condition, and is the admiration of all who
see it.”
R.M. Gorrie (1965), who was forester on the
Mansfield estate in 1919, stated that “the various
written accounts give 1860 as the date of planting,
but Lord Mansfield himself verified from records
that it was in fact 1858.” Booth (1890) mentions
1857 as the year of establishment. The
discrepancies between accounts probably stem
from the fact that authors did not distinguish
clearly between the Taymount plantation and a
13-acre plantation established in 1857 at Scone
(Hutchison 1879, McCorquodale 1880). The seed
source for these plantations were trees at
Lynedoch that Scone grew from the seed sent by
David Douglas in 1826.
The Taymount plantation was thinned in 1887
to 499 trees per ha. In 1896, all trees were pruned
to a height of 10 to 12 m (Somerville 1904).
Schlich (1888) estimated the volume of the
plantation to be
206.8 cm/ha at age 28 years. In 1900, at age 40,
to have slowed. Robinson (1914), after a 1913 follow- ing account of seed procurement from
survey of Douglas-fir plantations, states: “The 1920 to 1980:
Taymount plantation which by its volume The first seed imported by the Forestry Commission
production first drew general attention to the in 1921 was 913 lbs, reputed to be from Washington.
The following year 4,000 lbs came from the lower
Douglas fir, now proves to be the least
Fraser
vigorous of the woods examined.” According
to Gorrie (1965), data of standing timber were
not recorded after 1913, and no record exists
of the felling or disposal of the felled crop. He
believes that the trees of the Taymount
plantation were prob- ably cut about 1920,
shortly before the estate was sold by Lord
Mansfield in 1921. Matthews (1983), without
citing a source, gives 1923 as the year when
the trees of the Taymount plantation were
felled.
The Taymount plantation became a
showpiece of Douglas-fir, and this was an
important factor in extending the planting of
the species throughout Great Britain (Edwards
1957). Numerous planta- tions of Douglas-fir
had been established in the last quarter of the
19th century (Macdonald 1952), and
thousands of Douglas-firs had been planted as
soli- tary trees to fill vacant spots in existing
plantations (McCorquodale 1880).
Undoubtedly, many of the 19th-century
plantations resulted from home col- lections,
but not all. From 1870 to 1880, the firm of
C.H. Manning of Roy, Washington, supplied
seed, at first from near the lower Columbia
River and later from northwestern
Washington. Seed was also prob- ably
imported in the late 19th century from British
Columbia (Lines 1987). In general, these
early seed imports were made up of
provenances well suited for the British Isles,
although not always. Henry and Flood had
already written in 1920 that the interior
variety had been tested in numerous localities
but was invariably a failure. Kay and
Anderson (1928) also pointed out that several
plantations had been grown from unsuitable
seed, giving rise to poor- quality stands.
Cultivation of the species declined after the
turn of the century when the first enthusiasm
for the tree had waned. Creation of the
Forestry Commission in 1919, and the large
forestation program initiated by that
organization, led to renewed interest in the
planting of Douglas-fir. Lines (1987) gives the
Chapter 3. Areas of Introduction 37

River. Thereafter most seed up to the 1950s came states that Douglas-fir had begun to lose favor
from both these sources. Small amounts came from
the Shuswap Lake area in the interior of British
with private land owners, and Wood (1955)
Columbia, from Oregon or from undefined sources in mentions that the species lost favor in
‘USA’ or ‘BC’. During the next 30 years a few southeastern England because of its
seedlots came from Vancouver Island and Oregon, but
the bulk was from Washington. Of the total of 50,747 disappointing lack of vigor, adding “It is not at
lbs imported, 77 percent came from the USA and 23 all clear how far the troubles which have
percent from British Columbia. rendered it unpopular are specific, or how far
The predisposition of interior British Columbia race has entered into the matter.” A factor in the
prov- enances to Rhabdocline pseudotsugae has led decline in its use was the widespread
to them not being recommended for use in Great occurrence of infes- tation by the Cooley spruce
Britain (Alan Fletcher, personal communication). gall adelgid (Adelges [Gilletteella] cooleyi), which
Douglas-fir seems to have lost some of its appeal made the trees look sickly and slowed growth for
to woodland owners in the second quarter of the a period.4
A reversal of that trend took place in mid-
century. The data from the latest woodland
census (Locke
20th century. Scott (1931) noted a reduction in the
area planted to Douglas-fir on account of
4. Letter from Dr. D.C. Malcolm, University of Edinburgh, dated
prejudice against the wood of the species. 7 December 1993.
Macdonald (1952)

Table 3.1 Area stocked with Douglas-fir, according to census of woodlands in 1947 and 1982. Percentages are percent of all conifers
(from Edwards 1957, Locke 1987).

England Scotland Wales Great


Year ha % ha % ha % Britain %
ha
1947 Private woodlands 2917 4 2202 2 747 9 5866 3
Forestry commission 4251 6 2320 3 2861 9 9432 5
Total 7168 5 4522 2 3608 9 15298 4
1982 Private woodlands 12144 7 5190 2 3159 8 20493 4
Forestry commission 12919 6 6438 1 7549 6 26906 3
Total 25063 6 11628 2 10708 6 47399 4

Table 3.2 Area and standing volume of Douglas-fir in Great Britain by planting year classes (From Locke 1987).

All woodland ownerships


Planting year class Private woodlands (ha) Forestry commission Total (thousand m3 with bark)
Pre-1861 158 41 199 91.0
1861–1900 299 36 335 172.5
1901–1910 193 51 244 107.1
1911–1920 490 96 586 224.9
1921–1930 1260 2896 4156 1494.3
1931–1940 1031 1503 2534 773.3
1941–1950 1041 1786 2827 593.4
1951–1960 5116 9100 14216 1749.5
1961–1970 7499 8306 15805 879.5
1971–1980 3406 3091 6497 —
TOTAL 20493 26906 47399 6085.9
38 Douglas-fir: The Genus Pseudotsuga

1987) show that the area occupied by Douglas-fir Table 3.3 Standing volume of Douglas-fir in Great Britain by
increased between 1947 and 1982 by a total of size classes (from Locke 1987).
32,101 ha, of which 17,474 ha were on Forestry
Size class (dbh) Thousands m3 with bark Percent
Commission lands and 14,627 ha on private
7 to 20 cm 2525.0 41.5
woodlands. In 1982 Douglas-fir formed 3%
21 to 30 cm 984.4 16.2
(26,906 ha) of the total area of private coniferous
31 to 50 cm 1858.8 30.5
woodlands. The largest share of land stocked with
>50 cm 717.7 11.8
Douglas-fir was in England, followed by Scotland
Total 6085.9 100.0
and Wales (Table 3.1). A break- down of the area
in Douglas-fir by planting-year classes (Table 3.2)
shows a similar pattern of pe- riodic increases and ing growth of trees native to the Pacific Northwest
decreases in area planted to Douglas-fir for both in Scotland, prompting him to write, “Scotland is
Forestry Commission and pri- vate woodlands that incomparable in its wealth of immense conifers
appears to reflect changes in the popularity of the and is the Oregon/California of Europe.” He
species. Data for the area planted to Douglas-fir found a Douglas-fir in Craigvincan that measured
before 1900 do not truly represent that period. 60 m in height, which makes it probably the tallest
Macdonald’s (1952) remark, “Numerous fine Douglas- fir on record in Great Britain. That tree
plantations were created throughout the last was still growing at a rate of 0.3–0.5 m per year in
century such as the celebrated stand at Taymount 1993 (D.C. Malcolm, 1993, personal
in Perthshire, but most of them have now disap- communication).
peared,” indicates that more land was planted to The most favorable sites for growth are in the
Douglas-fir than shown by these data. By contrast, wet and moderately wet coastal areas on well-
the data for the years 1971 to 1980 reflect an drained loams and sandy loams of intermediate
actual trend of reduced planting of Douglas-fir fertility. But the sensitivity of Douglas-fir to
after a peak in the preceding decade. The area constant wind exposure to wind restricts the
occupied by the species has shrunk somewhat species to sheltered sites for best growth.
since the 1982 census because of its reduction Anderson (1961) considered Douglas-fir to be a
from 26,906 ha to 25,400 ha on Forestry species for the middle-hill slopes rather than the
Commission lands by 1987 (Lines 1987). valley bottoms and higher reaches. Cultivation of
However, the amount of land stocked with the tree is not necessarily confined to low
Douglas- fir on private holdings did not change. elevations, however. Douglas-fir has been grown
The standing volume of Douglas-fir in Great successfully at altitudes of 305 m in Scotland
Britain amounted in 1982 to slightly more than 6 (Macdonald 1952) and up to 408 m in Wales
million cubic meters. The 20- to 30-year, and 50- (Bennett and Long 1919), but because of the high
to 60-year-age classes contained the largest shares risk of top damage in wind-exposed areas, it is
of standing volume (Table 3.2). The largest best kept below 250 m.5
percentage of standing vol- ume was in pole-sized Silvicultural practices have influenced the distri-
timber, and the smallest in saw timber (Table 3.3). bution of Douglas-fir aside from environmental
Climate permits growth of Douglas-fir fac- tors. The species has been much used for
through- restocking poor scrub areas of oak and birch. As a
out much of the British Isles. Limits of result, the dis- tribution is to some extent
temperature and precipitation for growth of the governed by the situation of oak and birch scrub,
species in Great Britain are not clearly defined, which has been converted to high forest (Edwards
however (Wood 1955). Although the mild and 1957). Thus, in Scotland, the greatest
humid climate in the western parts of the country concentrations of sites stocked with Douglas-fir
provides highly favor- able growing conditions, are in Kirkcudbright, Argyll, Perth and Inverness
the species is also capable of making a because of the frequent use of Douglas-fir to
satisfactory but slower growing crop in the lower rehabilitate scrub lands (Anderson 1967).
rainfall and more continental climate
of eastern and southeastern England (Day 1955,
Streets 1962). Mitchell (1983a) reported outstand- 5. Ibid.
Chapter 3. Areas of Introduction 39
douglas-fir-612237.html.
Compared to two other western-American spe-
cies, Sitka spruce and lodgepole pine, which ac-
counted in 1982 for 40% and 10%, respectively,
of the total area stocked with conifers in Great
Britain (Locke 1987), Douglas-fir held a relatively
small share, 4%. That share may increase in the
future, however, according to Lines (1987): “In
view of its high timber value, rapid growth and
resistance to butt rot it is likely to be used on an
increasing scale as the emphasis swings away
from afforestation of bare ground to replanting of
more sheltered valley sites, e.g. those carrying
better quality Scots pine or larch.”

Ireland
Douglas-fir was introduced to Ireland shortly after
1850 (Fitzpatrick 1966). The first plantings
consisted of solitaires or small groups in arboreta
and selected spots in open woodland. Professor
Tom Clear wrote in 1951,
The results of this type of planting are to be seen in
many parts of Ireland, particularly in desmenes like
Powerscourt, Carton, etc. The growth of the
specimens thus planted at Kilruddery was most
remarkable and by the beginning of the present
century Douglas-fir was well on the way to
becoming a firm favourite in the race for pride of
place among the newer exotics. After seeing the
giants at Powerscourt and Carton one can well un-
derstand the superoptimism that prevailed with
regard to this species some 40 or 50 years ago.

The tallest Douglas-fir in Ireland stands at


Powerscourt Demesne, County Wicklow. Its
height was measured at 53.14 m in 1991, and its
age was about 125 years. By 2013, its height was
measured at
61.5 m, and it was officially recognized as not
only the tallest Douglas-fir, but the tallest tree in
Ireland since recordkeeping began.6
The use of Douglas-fir as a forest tree
coincided with the advent of State Forestry in
Ireland (Clear 1951). The species’ history of
planting is reflected by area occupied according to
age class (Table 3.4). Many of the early
plantations were on old woodland sites and
usually in mixture with Picea abies and Larix
europea. Planting of Douglas-fir ceased almost

6. Mr. Alistair Pfeifer, Forestry & Wildlife Service, Bray, Ireland,


letter of 16 February 1994; “Ireland’s tallest tree - a 200ft Douglas
Fir,” Breaking News, Ireland, 05/11/2013,
http://www.breakingnews. ie/ireland/irelands-tallest-tree-a-200ft-
transmission poles (O’Driscoll 1978). Expected
Table 3.4 Area occupied by Douglas-fir and mean yield rotations in state forests are from 40 to 60 years.
class in state plantations in Ireland by age class.

Age class Area (ha) Mean yield class


Pre 1920 9 14.4
1920 - 1929 165 14.4
1930 - 1939 208 13.7
1940 - 1949 23 12.4
1950 - 1959 483 13.2
1960 - 1969 2206 15.4
1970 - 1979 1874 17.7
1980 - 1989 2104 17.6*
1990 - 1993 701 17.5*
* Projection
Note: yield class = 1 m3/ha/y. Source: Data provided by A. Pfeifer, Irish
Forest and Wildlife Service.

completely in World War I and was not


resumed on any appreciable scale until 1921.
The early 1920s saw the establishment of
extensive monocultures with Douglas-fir on
sites favorable for the species. With further
extension of planting programs, site require-
ments of Douglas-fir received less attention.
As Clear (1951) phrased it “This departure
from sound selec- tion of sites could have but
one result-poor crops.” In addition,
appearance of the Cooley spruce gall adelgid
(Adelges [Gilletteella] cooleyi) and the Swiss
needle cast disease (Phaeocryptopus
gaeumannii) con- tributed to stagnation of
growth in plantations. All of these problems
led to a decline in popularity of Douglas-fir
and, by 1940, planting of it had virtu- ally
come to a halt. Moreover, existing stands
were regarded as being without future because
of their stagnant and debilitated appearance
(McEvoy 1943), and a policy of replacing
unsatisfactory stands of Douglas-fir was
instituted (Clear 1951).
Heavy thinnings during and after World War II
dramatically improved the remaining stands
and led Irish foresters to reassess the value of
Douglas-fir. The most notable plantations
saved by these rigor- ous stand openings are
in the Suir Valley, between Carrick and
Clonmel, and on Slievenamon, as well as in
the glens of County Wicklow (Fitzpatrick
1966). The interest in Douglas-fir to Irish
forestry lies in the high returns that may be
obtained in the lucrative market for
40 Douglas-fir: The Genus Pseudotsuga
ous publications and in the literature is usually referred to as John
Douglas-fir does well on sheltered sites with Booth.
deep, moist, and well-drained soil. It grows poorly
on exposed sites, on peat, and on lime soils, all
conditions common in Ireland. Thus, the species
is restricted, to the lower slopes of the mountains
in Counties Louth, Tipperary, Waterford,
Wexford, and Wicklow (Fitzpatrick 1966).
Only the coastal variety of Douglas-fir is
suitable for Irish conditions. Much of the seed
imported in the past apparently represented
provenances from the coastal regions of
Washington. Early results from Irish
participation in the IUFRO international
provenance trials indicate that coastal
provenances from Washington hold considerable
promise for use in Ireland (O’Driscoll 1978,
Pfeifer 1988).
The 1958 woodland census in the Irish
Republic recorded nearly 1,375 ha of pure stands
of Douglas- fir with a volume of 274,096 m 3, or
an average of 199 m3/ha, the highest for any
species in Ireland (Fitzpatrick 1966). In the 1968
woodland census (O’Flanagan 1968), Douglas-fir
represented only 3% of the total afforested area,
and had a yield class range from 2 to 24, with the
weighted average of 14. In 1993, Douglas-fir
occupied 7,772 ha in the Irish State Forests. The
area occupied by Douglas-fir plan- tations in the
private sector is very small, amounting to about
500 ha.7 That Douglas-fir assumes a minor role in
the species composition of Irish forests re- flects
the fact that its planting range is limited by the
availability of suitable sites. Although Douglas-fir
is regarded as the premier conifer in Ireland, many
foresters shy away from it because of difficulties
with its establishment, and they plant Sitka spruce
(Picea sitchensis) instead.8

Western Central Europe


Germany
Douglas-fir was introduced to Germany shortly
after it had come to Great Britain. John Richmond
Booth (1800-1847), a nurseryman in Flottbeck
near Hamburg, was a member of the Royal
Horticultural

7. Ibid.
8. Ibid.
9. John Cornelius Booth did not use his middle name in his numer-
Society. He received some of the Douglas-fir Brandenburg Forestry Association (Märkischer
seed collected by David Douglas when it was Forstverein) at Neubrandenburg, marking the
distributed to members of the Society beginning of wider interest in the species. That
(Kremser 1974). In 1829, Booth planted a 2- Douglas-fir commission, consisting of six men
year-old seedling raised from the seed with Count Willamowitz- Moellendorf as
collected by David Douglas in his arboretum, chairman, began its work by dis-
which was then the largest in Europe, next to
that of the Duke of Bedford at Woburn
Abbey. That seedling was probably the first
Douglas-fir ever planted in Germany (Booth,
1877). It developed into a tree that was felled
in 1882 by John Cornelius Booth,9 son of
John Richmond, to demonstrate the quality
of Douglas-fir wood grown in Germany (J.
Booth 1882). The elder Booth was a strong
advocate for intro- ducing of North American
tree species to Germany. In an address to the
Society of German Agronomists and
Foresters at its meeting in 1841 at Doberan,
Mecklenburg, he told his audience that
Douglas- fir grows well under the climatic
conditions of Germany, and recommended
trials with that spe- cies (J.R. Booth 1841, p.
51-52). J.R. Booth’s efforts were not entirely
without success. An inventory of foreign tree
species planted in Germany before 1880
(Weise 1882) recorded Douglas-firs in many
parts of Germany, although their numbers
were very small. In 1988, the oldest still-
standing Douglas-firs, one at Jaegerhof in
Pommerania, the other at the
Barneführerholz near Oldenburg, had come
from the Booth Nursery in Klein-Flottbeck.
The Jaegerhof tree had been planted as a 4-
year-old seedling in 1842. The planting date
for the Barneführerholz tree is 1843, but
information is lacking about its age at time of
planting (Rothkirch and Struthoff 1989). J.
Booth (1907a) emphasized that both trees are
of the
«green» variety and had been grown from
seed that
had come from the American Northwest. That
the seed had been part of the original David
Douglas collection, as suggested by Rothkirch
and Struthoff (1988), is unlikely because 11-
year-old seed would scarcely have been viable
without the availability of cold storage
facilities.
A «Kommission für die Douglas-fichte»
was es- tablished at the 1878 meeting of the
Chapter 3. Areas of Introduction 41

tributing Douglas-firs to interested forest owners menziesii established 1889, 1906, 1914, and 1916
in spring of 1879. The seed was supplied by J. in the Kaliningrad (formerly Königsberg, East-
Booth, and came from the northernmost part of Prussia) region of Russia (Fedorov 1981). In 1981,
the range of Douglas-fir (Booth 1882, p. 34) the oldest of these plantations had a standing
Booth (1880) delivered a paper on exotic tree volume of 1,160 m3/ha. Trees in the plantations
species at the meeting of the Union of German are hardy and flower every 3 or 4 years,
Forest Experiment Stations in September 1880 in producing seed of good quality. Most of the seed
Baden-Baden at the request of the Prussian Forest used in the Prussian State Forests before 1890 was
Experiment Station. In that paper, Booth outlined probably supplied by J. Booth. From 1891 to
the advantages to be gained by introducing exotic 1895, the U.S. Bureau of Forestry shipped about
tree species into the forests of Germany. His 145 kg of Douglas-fir seed to Prussia. The origin
presentation must have been convincing because of that seed is unknown. Beginning in 1896,
the members of the Union decided to initiate purchase of seed was at the discretion of State
systematic trials with North American and East Forest Districts, who usually bought it from the
Asian broadleaves and conifers. The working plan least ex- pensive sources. Since about 1909, seed
for these trials, which was agreed upon by the procurement for the Prussian State Forests was
members of the Union at their meeting in August handled by the German Dendrological Society. The
1881 at Braunschweig, in- cluded Douglas-fir Society collected seed of the var. menziesii
among the conifers selected for the trials primarily in the Cascade Range of Washington
(Ganghofer 1884). and Oregon (Kanzow 1937). The Brunswick
The testing program of the Prussian Forest Forest Experiment Station estab- lished small trial
Experiment Station spanned an area from the Eifel plantations in 1876 with 3-year-old Douglas-fir
Mountains in the west to East-Prussia in the east. seedlings. Puchert (1954) presumed that the seed
That provided the opportunity to observe the from which these seedlings were raised came
performance of Douglas-fir under climates from British mother trees established from seed
ranging from mild oce- anic to harsh continental. brought by David Douglas. In 1880, the
Bernhard Danckelmann, head of the Prussian Experiment Station bought seed from three firms
Forest Experiment Station, reported in 1884 that — Appel, Nungesser, and Trumpff—that
plots of Douglas-fir totaled 87 ha; by 1890, that represented unsuitable provenances and led to
number had expanded to nearly 140 ha plantation fail- ures. By contrast, seed obtained
(Schwappach 1891). Danckelmann’s succes- sor, from J. Booth in the years 1881 to 1888 gave such
Adam Schwappach, already concluded after the good results that the Experiment Station
first decade of trials in 1891 that Douglas-fir encouraged operational planting of Douglas-fir as
should be introduced into the Prussian forests. He early as 1899 (Puchert 1954). The origin of the
reiterated that conclusion in subsequent reports seed supplied by Booth is unknown, as is that of
(1901, 1911) and stated in his final report (1920) seed lots obtained from a dozen different seed
that Douglas-fir was particularly suited for the dealers from 1886 to 1910 (Grundner 1921). After
sandy soils of northern Germany, where Norway World War II, Brunswick became part of the state
spruce (Picea abies) is a common, but off-site of Lower Saxony, which in 1950 had 1,850 ha of
tree. He was not alone in his assessment. Already Douglas-fir (Borchers 1951). From 1950 to 1992,
in 1904, at the meeting of the Society of East- and the acreage of Douglas-fir in Lower Saxony
West-Prussian Foresters, participants had increased
expressed the opinion that Douglas-fir was to about 15,000 ha.10
suitable for use in East-Prussia (Anonymous The first plantings of Douglas-fir in the then
1904). Additional accounts of perfor- mance of king- dom of Saxony were made in 1878 in the
Douglas-fir were published by Reichenau (1911) Tharandt District of the Saxon State Forests, and
for West-Prussia, and by Böhm (1922) for East- in the Plauen City Forest (Zacharias 1931).
Plantations of the spe-
Prussia. These early assessments of the suitability
of Douglas-fir for use in East-Prussia were shown
10. Dr. H.J. Otto, Lower Saxon Ministry of Nutrition, Agriculture
to be correct by a report on 4 plantations of P. var. and Forestry, letter dated 29 July 1993.
42 Douglas-fir: The Genus Pseudotsuga
cies were established in state and municipal 11. Dr. H. Weissgerber, Hesse Forest Experiment Station, letter 10
forests in the next 25 years on a very small scale, March 1980.
with seedlings of the variety menziesii. Most of
the seed appears to have been supplied by J.
Booth. The initial results were sufficiently
encouraging that Nobbe (1895) recommended
wider use of Douglas-fir. But not until 1904 did
the Saxon Forest Experiment Station begin
systematic trials with both the coastal and interior
va- rieties of the species. Neger (1914), who
reported on trials located at elevations of 100 m in
the lowlands to 1,000 m in the Erzgebirge (Ore
Mountains), stated that the coastal variety had
performed well in the lowlands and the interior
variety had grown better at high elevations. A
detailed review of the distribu- tion of Douglas-fir
after World War I in Saxony by Zacharias (1931)
showed that of the 87 State Forest Districts, 47
contained Douglas-fir in 1928. The area occupied
by stands of Douglas-fir was small, namely 88 ha
or 0.05% of all state-owned forests. Apparently, the
area occupied by Douglas-fir did not increase
greatly in the next 50 years. M. Hartig (1980)
listed 1,394 ha stocked with Douglas-fir for a
region that covered the major part of the former
state of Saxony. The first plantings of Douglas-fir
in the State of Hesse were made in 1858 in the
Darmstadt City Forest (Walther 1911), although
systematic trials with the species did not begin
until 1884. With few excep- tions, like the
Büdingen City Forest where 70 ha were planted to
Douglas-fir from 1898 to 1917 (Spengler 1925),
introduction of the species proceeded rather
slowly. By 1945, only 722 ha of Douglas-fir
planta- tions existed. Then, the pace of planting
Douglas- fir accelerated. By 1967, about 5,000 ha,
or 0.6% of the total forest area of Hesse were
stocked with Douglas-fir (Groos 1968). By 1980,
the acreage of Douglas-fir had increased to about
15,000 ha (37,050 acres) (Riebeling 1979,
Weissgerber 198011). Based on percentage, the
increase was largest in private and communal
ownerships where the share of Douglas- fir
constituted 2.3% in each of these two ownership
categories, but it was only 1.2% in the State
Forests. Douglas-firs up to 20 years of age
existed in Bavaria by 1880, but they were too
few to permit
any conclusions about their value as forest in the former grand duchy of Baden date back to
trees (Ganghofer 1884). The Bavarian Forest 1860 in the forest district of Oberweiler, and to
Experiment Station began with systematic 1865 in the city forests of Heidelberg and
trials of Douglas-fir in the decade 1881 to Freiburg. By 1880, Douglas-fir had been planted
1891. These first trials were conducted on a in 20 locali- ties (Wimmer 1909). The early trials
small scale, that is, planting of no more than between 1870
50,000 trees. Seed used in these trials was
obtained through J. Booth; Prof. C.S. Sargent;
and the seed dealers Robert Douglas in
Waukegan, Wisconsin; Keller in Darmstadt;
and Steingaesser in Miltenberg (R. Hartig
1892).
Encouraged by initial success, nearly one
mil- lion trees were planted in State Forests
from 1891 to 1904 (Mayr 1907). A
questionnaire sent to 156 forest districts in
1905 about the performance of Douglas- fir
provided mostly positive responses, and led to
increased planting of Douglas-fir until 1913.
Planting of Douglas-fir nearly ceased during
the next 10 years because seed imports
stopped during World War I and during the
first 5 years after the war. A census of foreign
trees by the Bavarian Forest Service in 1923
showed that 157 ha were stocked with
Douglas-fir. The inventory was incomplete,
but the total area oc- cupied by Douglas-fir
was unlikely to be more than 200 ha,
indicating considerable losses of Douglas-fir
plantations. Harrer (1925) attributed these
failures to disregard for provenance and
insufficient knowledge of the silvical
characteristics of the species.
The next census of Douglas-fir in the Bavarian
State Forests was undertaken almost half a
cen- tury later in 1969 (Bayerische
Staatsministerium für Ernährung,
Landwirtschaft und Forsten 1970). The results
showed the existence of 1,248 ha of Douglas-
fir plantations, of which 66% had been
established after 1950. Since 1970, the
acreage of Douglas-fir in Bavaria has
increased considerably. According to a
questionnaire addressed to forest owners, the
annual area planted to Douglas-fir averaged
600 ha (Huss and Siebert 1976). Först (1980)
mentions 2,000 ha of Douglas-fir stands just
in Lower Franconia in 1980, but figures for all
of Bavaria in 1980 or afterwards are lacking
in the literature.
The oldest known plantings of Douglas-fir
Chapter 3. Areas of Introduction 43
forest area of that region.
and 1880 were largely failures because of frost
kill (Mörmann 1956). Although Baden was
among the first German states that participated in
the trials initiated by the German Union of Forest
Experiment Stations in 1880, the Grandducal
Forest Directorate issued a directive in 1884 that
limited the cultivation of Douglas-fir to forest
districts designated as trial participants. That
directive remained in force until 1899 (Klumpp
and Gürth 1988). Results of the trials begun after
1880 were summarized in a report by Alten
(1898) that was distributed to all Baden State
Forest Districts. His favorable assessment of the
performance of Douglas-fir provided the impetus
for increased cultivation of Douglas-fir in Baden.
A total of 53 kg of Douglas-fir seed had been
distributed to State Forest Districts in Baden from
1883-1899 (Wimmer 1909). The seed,
presumably of Washington and Oregon origin,
was supplied by the firms of J. Booth and Sons
and G. J. Steingässer. Between 1898 and 1910,
seed was obtained from seed dealers in Victoria,
British Columbia, and Roy, Washington (Klumpp
and Gürth 1988). Plantations established between
1880 and 1910 developed into stands of high
quality. By some fortunate circum- stance, seed
from which these trees were grown came from
provenances well suited for this part of Germany.
Later imports of seed, especially those directly
from the Long Bell Company in Washington in
the years 1927-1935, often contained unsuitable
provenances and resulted in poorly growing
stands (Mörmann 1956b).
In spite of the efforts of many foresters to plant
Douglas-fir (Oeschger 1975), the area occupied by
the species in publicly owned forests remained
small for decades. Baden had 112 ha, stocked
with Douglas-fir in 1906 (Wimmer 1909), 150 ha
in 1920 (Hausrath 1921), 505 ha in 1930 (Killius
1931). After World War II, the pace of planting
Douglas-fir quickened, and in 1960 about 4,000
ha were stocked with Douglas-fir in public forests
(Scheifele 1965). Douglas-fir stands are not
evenly distributed throughout Baden but are
concentrated in the Odenwald and the west slope
of the Black Forest (Behler 1980, Weidenbach
1980). Public forests in the Upper-Rhine region
already contained 7,065 ha of Douglas-fir in 1975
(Schülli 1986). That amounts to 8% of the total
from lands be- longing before the war to Bavaria
Special mention must be made of the and Prussia, has perhaps the largest concentration
Freiburg City Forest. Here, Douglas-fir has of Douglas-fir stands in Germany. In 1986, the
been planted since 1901. In 1948, the Freiburg species occupied 29,400 ha in both private and
City Forest contained about 220 ha (543 public ownerships, or
acres) stocked with Douglas-fir. At that time,
that was probably the largest area of Douglas-
fir in a single forest in Germany (Seibert
1951). The Freiburg City Forest and the
nearby Kandern and Sulzburg State Forest
Districts belong to the best Douglas-fir areas
in Germany. Nearly all are in site class I and
stands have a mean annual increment at age
100 of 15-18 m3 (Volk 1959).
In Wuerttemberg, Prof. Lorey, Head of the
Royal Wuerttemberg Forest Experiment
Station, initiated the first trials with Douglas-
fir in the years 1882 to 1892. By 1890, 21 trial
plots with a total area of 3.87 ha had been
established (Lorey 1890). In a follow-up
report to his 1890 paper, Lorey (1897)
indicated that between 1891 and 1895, 62 kg
of seed of the coastal variety, and 10 kg of
seed of the interior variety, had been
distributed to State Forest Districts by the
Royal Forest Directorate. In 1896, Douglas-fir
was represented in 18 State Forest Districts.
An additional 257 kg of seed of the variety
menziesii was distributed until 1908, and by
1911, 67 ha of Douglas-fir planta- tions had
been established (Holland 1912). The next
four decades saw only a small increase in the
share of Douglas-fir in the state forests of
Württemberg. In 1950, only 368 ha were
stocked with Douglas-fir (Zimmerle 1952).
Following World War II, Baden and Württemberg
were combined into one state, and recent
statistical data on the area stocked with
Douglas-fir do not provide a breakdown
between the former two states. Between 1960
and 1970, planting of Douglas-fir in- creased
considerably. By the end of the decade, the
area occupied by Douglas-fir was assumed to
be in excess of 10,000 ha (Volk 1969). By
1984, the area stocked with Douglas-fir had
grown to about 20,000 ha. The annual harvest
of Douglas-fir in publicly owned forests in
Baden-Wuerttemberg amounted to about
25,000 m3.
The State of Rhineland-Palatinate
(Rheinland- Pfalz) created after World War II
44 Douglas-fir: The Genus Pseudotsuga
6% of the total forest area of Rhineland-Palatinate excessive risks, and Douglas-fir planting was resumed
(Petri 1986). in the 1950s on a scale larger than before.
Beginning as early as 1865, some private forest
owners also began to introduce Douglas-fir into
their forests with notable success. The most
famous are the Douglas-fir plantations in the
Sachsenwald estate of Prince Bismarck in
Schleswig-Holstein (Titze 1906, 1920), and the
former estate of Count Wilamowitz in the northern
Elbe River region (Wilamowitz- Möllendorf
1907, Zeidler 1956). The seed, of the coastal
variety of Douglas-fir, was supplied by the firm of
J. Booth and Sons. The area of Douglas-fir at
Gadow increased from 17 ha in 1920 to 131 ha in
1949. Since 1922, only seed from Douglas-fir
mother trees at Gadow was used for establishing
new plan- tations (Adolph 1936, 1949).
Foremost among the advocates of Douglas-fir
was J. Booth, who worked for the introduction of
the species with the zeal of a crusader (1877,
1880, 1882, 1890, 1896, 1903, 1904, 1907a,b) and
later Carl
Alwin Schenck (1928, 1939). But not everybody
shared that enthusiasm for introducing Douglas-
fir. Reuss (1885) cautioned that a valid judgement
of the species’ merits in Germany would be
possible only after 100 to 150 years of experience.
Boden (1902) argued that Douglas-fir was
actually inferior to na- tive European conifers,
typical of an opinion appar- ently held by a few
practicing foresters in Germany at that time.
However, few seem to have gone as far as a
“Forstmeister” who purposely left Douglas-fir
raised by his predecessor in the nursery beds, and
after a few years, simply disposed of the young
trees as too old for transplanting (Mayr 1907).
Planting of Douglas-fir declined sharply from
1914–1924 because World War I, and the ensuing
period of monetary instability in Germany inter-
rupted the import of seeds from North America. A
further setback to cultivating Douglas-fir occurred
when the rapid spread of the Swiss needle cast
patho- gen (Phaeocryptopus gaeumannii) assumed
epidemic proportions in the late 1930s. As a result,
planting of Douglas-fir was discouraged or
outright prohibited, as in Württemberg in 1940
(Merkle 1951). Increasing knowledge of the
biology of the fungus showed the conditions
under which Douglas-fir could be grown without
The Federal Republic of Germany had about Phaeocryptopus gaeumannii soon dis- couraged
6,000 ha of Douglas-fir in 1950 and, based on further cultivation of the interior variety. A
Knell’s (1960) estimate of an annual rate of directive issued by the Bavarian Forest Service in
increase of 1,250 ha per year, had about 1932 led to the systematic removal of trees of the
31,000 ha stocked with the species by 1970. interior variety in existing stands (Foerst 1980).
In 1987, the Federal Republic of Germany,
(excluding the German Democratic Republic,
had about 80,000 ha of Douglas-fir. Forest
planning pro- grams allotted to Douglas-fir
10% to 20% of the total forested area in the
Federal Republic of Germany. Statistics on
Douglas-fir covering all of the former German
Democratic Republic (GDR) have not been
published. According to a letter of August 10,
1970 by, Landforstmeister Vonhof, Douglas-
fir occupied between 10,000 and 14,000 ha, or
about 0.5% of the total forest area of the
GDR, at that time. Apparently, the area
stocked with Douglas-fir did not increase
substantially in subsequent years. Nearly 20
years later, Braun and Weissleder (1986)
pointed out that the goal was to increase the
area occupied by Douglas-fir in the lowlands
of the GDR from less than 1% to 7.6%, and,
in the highlands, from 0.16% to about 3%.
Based on the above data, the assump- tion that
the area of Douglas-fir in 1990 approached
100,000 ha in the united Germany seems
justified.
Nearly all of the Douglas-fir in Germany
belongs to the coastal variety of the species.
Stands estab- lished in the last quarter of the
19th century with seed of the variety
menziesii showed exceptional growth, and
prompted repeated attempts in the second half
of the 20th century to trace the origin of these
early seed imports. All such efforts remained
unsuccess- ful, for two reasons: (1)
designation of seed origin was not customary
when seed was shipped and (2) records that
may have provided relevant informa- tion are
long lost.
Seed of the variety glauca came first to
Germany between 1891 and 1895 (Heyder
1986). Its origin is also unknown. Seed from
interior British Columbia, designated as
variety caesia in Germany, was im- ported
regularly from 1902 to 1912 (Fürstenberg
1923). The generally poor growth of the
inland form of Douglas-fir, and its
susceptibility to Rhabdocline pseudotsugae and
Chapter 3. Areas of Introduction 45

A recommendation was made again in 1950 to


for agricultural purposes. Booth (1907a) cites an
eliminate all stands of interior Douglas-fir by cut-
article in the November 6, 1895 issue of Garden
ting them as soon as they start to flower (Schippel
and Forest, which indicated that the largest
1950). The experience with interior Douglas-fir,
Douglas-fir in the 47 years since planting had
however, has not been negative everywhere.
reached a height of 18.5 m and a dbh of 49 cm, a
Hartig (1997) assessed the performance of coastal
remarkable perfor- mance on poor dune sand.
and in- terior Douglas-fir in three stands in the
Hacke-Oudemans and Oudemans (1955)
Saxon Ore Mountains (Sächsische Erzgebirge).
attempted to trace the origin of Douglas-firs
These stands had been established in 1905, 1909,
planted before 1870. They concluded that these
and 1936, with var. menziesii and var. glauca
trees apparently came from three different
planting stock. Trees that belonged to var.
sources. Some were progeny of the original David
menziesii had succumbed to frost injury soon after
Douglas trees in Scotland, but others were raised
planting, and those that be- longed to the southern
from seed obtained from Canada. A document in
subgroup of var. glauca died after infection with
the Gelderland Archives states “Through Mr.
needle cast fungi. By contrast, growth of the trees
Eduard Hamp of Victoria, British Columbia, seed
that belonged to the northern subgroup of var.
has been shipped to the Dutch Government from
glauca was equal or better than that of Norway
Douglas- firs 300 feet tall and over 9 feet in
spruce, and their tolerance of SO 2 emissions was
diameter.” The seed from which the trees of the
far superior to that of spruce. Cone crops of these
Schober plantation were raised stemmed from
Douglas-firs provided enough seed to permit
another but unknown location in northwestern
establishing several small plantations totaling
America. The oldest, still existing Douglas-firs,
about 25 ha from 1960 to 1984. This second
planted in 1857 (Veen 1951) are in the park of the
generation progeny grew vigorously and showed
royal palace “Het Loo.” The tree had been used in
resistance to frost and SO 2 injury.
plantations, but with varying degrees of success.
Growth of Douglas-fir, with mean annual
In 1899, the Dutch “Heidemaatschappij” (soci-
incre- ment (m.a.i.) of 100 years as the
ety for the reclamation of uncultivated land) in-
criterion, decreases from the south to the north,
stituted the “Committee for the Study of Exotic
and from the west to the east of Germany (Jahn
Coniferous Trees.” To learn more about the
1959). The best growth is in the Black Forest
potential of Douglas-fir for Dutch forestry, the
region of southwestern Germany and the
committee commissioned a comprehensive study
poorest growth is in the diluvial plains of
of Douglas-fir in the Netherlands that covered 29
northeastern Germany. But even where Douglas-
stands ranging in age from 19 to 67 years. Results
fir is doing poorly, its growth is often better
of the investiga- tion (De Hoogh 1924) led to the
than that of native species on the same site.
conclusion that Douglas-fir would be of
The Netherlands considerable advantage to Dutch silviculture, but
The indigenous forests of the Netherlands are that more knowledge of the species was needed,
mix- tures of oak (Quercus robur) and birch especially about the question of provenance.
(Betula spp.) and, on the better sites, oak and Probably on the basis of that report, a Douglas-fir
hornbeam. The financial yield from the native provenance study was begun in Holland as early
forest is low, and thus considerable efforts have as 1923.
been made to obtain higher returns by introducing The area occupied by Douglas-fir in Dutch
conifers (Van Soest 1956). Of the exotic conifers forests remained fairly small until the 1940s. At
used in Dutch silvicul- ture, Douglas-fir has long the end of World War II, several hundred hectares
been considered the most important. were stocked with Douglas-fir (Fovernied 1946).
One of the earliest plantings of Douglas-fir in Five years later, Veen (1951) estimated that
the Netherlands dates to 1848, when J.H. Schober Douglas-fir occupied about 4,000 ha. The share of
of Amsterdam established a plantation of exotics the species increased between 1939 and 1959 to
that included Douglas-fir on dune land unsuitable nearly 30% of the plantations in the 1- to 40-year
age classes (Van Soest 1959). The 1969 census
lumped Pseudotsuga and
46 Douglas-fir: The Genus Pseudotsuga
Larix together, and listed the area occupied as because the use of more suitable provenances has
32,000 ha. The area was about evenly divided eliminated the problem.15
between the two species (Wolterson, personal
communication). Results of the IUFRO Douglas-
fir provenance study, begun in 1967 (Kriek 1974),
will help in the selection of provenances best
suited to growing conditions in the Netherlands.
Of a total forested area of 342,000 ha in the
country, some 18,399 ha
were stocked in 1996 with Douglas-fir.

Belgium
The planting of 50 seedlings of Douglas-fir in
1872 on the estate of Count Visart at Sibret near
Bastogne in the Ardennes (Visart and Bommer
1909) was probably the first introduction of the
species into Belgium. The results were so
encouraging that Count Visart planted another
30,000 Douglas-firs between 1878 and 1909. The
interior variety was also tested at Sibret but grew
very poorly. Unfortunately, the Douglas-firs of
Sibret were destroyed in World War I (Hickel
1922). Another early introduction is a grove of
Douglas-firs planted about 1880 on a private
estate in Antwerp Province (Geelhand 1954). The
species was not widely planted during the next 50
years, however. The age-class distribution of
Douglas-fir stands in Belgium, listed in the 1958
general census of Belgian forests, indicated that
Douglas-fir occu- pied about 200 ha in 1930
(Ministère de l’Agriculture 1958). By 1950, the
area covered by Douglas-fir had increased to
about 1,500 ha (Gathy 1956) and, in the next 10
years, doubled to nearly 3,100 ha.12 The area
stocked with Douglas-fir continued to grow, and
in 1970, had attained 7,200 ha, amounting to 1.2%
of the total forested area in Belgium (Nanson
1978). According to an estimate by Nanson, 13 the
species currently occupies between 10,000 and
15,000 ha.
Climate in all parts of Belgium, from the coast
to the high plateaus of the Ardennes, permits
growth of Douglas-fir. The species grows best on
soils with good drainage, such as the “sols bruns
acides.”14 The Ardennes offer the best conditions
for cultivation of Douglas-fir (Galoux 1952).
Rates of failure used to be rather high in
plantations above 500 m (1,640 feet) elevation
(Delvaux 1964). But that is no longer the case
Only the variety menziesii is suitable for Belgium, letter of 22 July, 1993.
14. Corresponds approximately to inceptisols.
planting in Belgian forests. Washington
15. Nanson, Station de Recherches Forestires, Gembloux, Belgium,
provenances from the region between letter of 22 July 1993.
Darrington and Hoquiam are best for
planting in Belgium (Nanson 1978). Today,
Belgian foresters consider Douglas-fir to be
the most promising species for timber
production in Belgium, which is
understandable in view of its performance. A
stand at Mésy (Ardennes) produced 1,297
m3/ha that includes volume removed by
thinning, and a mai of 23 m3/ha (Poncelet
1963). These values rank among the highest
documented. The growth rates of the
species compare very favorably with those
of other productive species. At the end of
the rota- tion, which varies from 60 to 80
years, production of Douglas-fir stands
varies between 12 and 24 m3/ha/ year
depending on provenance and site quality.
To put that in proper perspective, average
production of Belgian forests amounts to
about 5 m3/ha/year, while that of Norway
spruce, the most profitable species in the
past, averages 12 to 13 m3/ha/year (Nanson
1978).
The trend toward rising importance of Douglas-
fir in Belgian forestry continues. That is
reflected by the fact that Douglas-fir
represents 20 to 25% of the species utilized in
the current annual planta- tion establishment
in Belgium ranging from 5,000 to 10,000 ha.

Luxembourg
Douglas-fir was introduced to Luxembourg in
1850, when it was planted at the “Jardin
Linden” in Limpertsberg (Modert 1965). A
tree planted in 1865 in the inner courtyard of
Meysemburg castle was, in 1965, the oldest
existing Douglas-fir in Luxembourg. At that
time, the tree was 32 m high and had a dbh
Soef v2e9rcaml s.mall stands of Douglas-fir
estab - lished in subsequent years near Meysemburg were
cut during World War II, and no records exist
about their performance. A Douglas-fir
plantation from the year 1883 in the
communal forest of Grevenmacher

12. J. Delvaux, Station de Recherches des Eaux et


Forêts, Groenendaal, Belgium, letter of 11 May
1970.
13. A. Nanson, Station de Recherches Forestieres, Gembloux,
Chapter 3. Areas of Introduction 47

was the oldest existing stand in 1965. It was a 6°00’ 6°30’

rem- nant of a much larger stand, which was so


BELGIUM Douglas-fir stand
heavily damaged during World War II that a large
portion of it had to be cut (Decker 1965). 05 10 Km

The state acquired an abandoned farm in 1902


and began experimental planting of exotic
50°00’ 50°00’
conifers, in- cluding both coastal and interior GERMANY
Douglas-fir (Modert 1965). The Douglas-fir seed
was obtained from the firm of J. Booth in
Hamburg, Germany. Data are not available on the
LUXEMBOURG
area planted to Douglas-fir in the following years;
all that remains is a statement that World War I
caused a reduction in planting of the species and
that planting ceased completely in the 1930s
(Decker 1965).
Interest in the species was apparently renewed Luxembourg

after World War II. Because records on origin and


performance of Douglas-fir in Luxembourg were
non- existent, Decker (1965) made an inventory, 49°30’ 49°30’

and took growth measurements of all the FRANCE


Douglas-fir stands older than 30 years (Figure
3.2) in Luxembourg. He concluded that Douglas-
fir is probably the most de-
sirable exotic species for cultivation in he planted several trees because Hickel mentioned
Luxembourg. Particularly on south and southwest that the last of those planted in 1842 was still alive in
exposures, the species grows better than Norway 1922.
spruce. Although the seed origin of the stands
inventoried by Decker is unknown, he presumed
that coastal provenances from western
Washington are probably best for use in
Luxembourg. The area stocked with Douglas-fir
in 1991 amounted to 1,674 ha, according to
Edmond Lies, Directeur des Eaux et Forêts of the
Grand Duchy of Luxembourg, as cited by De
Champs (1997).
France
Two dates appear in the literature for the intro-
duction of Douglas-fir to France. The statement of
Fourchy (1954) that the species was brought to
France in 1827 is questionable because no records
are known of distribution of the original Douglas
seed by the Royal Horticultural Society to
someone in France. More likely in accordance
with the facts is the ac- count by E. A. Carrière,
cited by Hickel (1922), that Monsieur Gautier-
Lachase made the first planting of Douglas-fir in
1842 at a place called Pré de l’Aulne (alder
meadow) in the vicinity of Louvigné-le Désert
(Dept. Ille-et-Vilaine) in the Bretagne. Apparently,
Figure 3.2 Douglas-fir stands in Luxembourg (from Decker 1965).

In 1844, shortly after that first planting, the


Marquis de Vibraye introduced Douglas-fir
into his park at Cheverny (Dept. Loir-et-
Cher). Carrière reported that these trees have
produced viable seed since 1851. Another of
these early introductions is a group of 30
Douglas-firs planted between 1842 and 1844
at Bord, the property of the Vicomtesse de
Sèze, near the village of St. Priest-Taurion
(Dept. Haut- Vienne). Slightly more recent
are the plantations in the Park of Harcourt
(Dept. Eure), belonging to the National
Academy of Agriculture, made from 1852 on
by Monsieur Pépin.
David Cannon began to plant Douglas-fir
about 1875 at his estate at Vaux not far from
Salbris in the Sologne. He was so impressed
with the performance of the species that he
wrote in 1909, “I believe that Douglas-fir is
by now the best acclimatized of exotic
conifers everywhere it has been planted.”16
At first, plantings were made mostly in
parks and along highways. Only since 1890
has Douglas-fir been considered for planting
in forests. These early

16. Translated from the original: L’Abies Douglasii est, jusqu’à


présent et partout, je crois, ou il a été planté, le triomphe de
l’acclimatation des conifères exotiques” (Cannon 1909).
48 Douglas-fir: The Genus Pseudotsuga
forest plantations are mainly in the Beaujolais.17 forestation, a development that was instrumental
Although small plantations of Douglas-fir existed in bringing about the rapid growth of the area
in at least some 30 of France’s 97 departments by stocked with Douglas-fir. By 1956, the area had
1920, the total area occupied by the species had increased to more than 10,000 ha (Pardé 1956),
remained small because of the lack of interest in and by 1970, was estimated to have risen to
Douglas-fir by the country’s Forest between 100,000 and 150,000 ha. Data collected
Administration (Hickel 1922). by the French National Forest Inventory showed
World War I, during which about 650,000 ha 220,000 ha of Douglas-fir stands (Bouchon 1984).
of French forests were destroyed, temporarily in- In 1993, the area stocked with Douglas-fir had
terrupted reforestation activities. Several hundred grown to 333,000 ha, an area larger than
kilograms of Douglas-fir seed shipped to France anywhere else in Europe (De Champs 1997a). The
immediately after the war provided some of the greatest concentration of the species is in the
planting stock for rehabilitation of these forest Massif Central including the Morvan.
lands (Podhorsky 1927). These seed shipments Considerable areas are also occupied by Douglas-
may well have consisted of unsuitable fir in the north- east and northwest of the country,
provenances because few Douglas-firs are present but the species is sparsely represented in the
in the stands planted right after World War I.18 southeast and southwest of France (Figure 3.3).
Notwithstanding seed problems, the area planted Of the lands stocked with Douglas-fir, 82% are
to Douglas-fir gradually increased between the private and only 18% public.
two world wars, especially in the western part of About 312,000 ha—that is, nearly 94% of the total
the Massif Central (Limousin, Plateau des area stocked with Douglas-fir in 1993, contained
Millevaches), but also in other parts of France. By stands younger than 35 years. That age-class dis-
1937, the area occupied by the species had grown tribution reflects the annual rate of about 10,000
to about 4,000 ha. A detailed account of the ha for planting of Douglas-fir from 1960 to 1980.
locations where these plantations had been That rate has markedly decreased since 1980, as
established and their size is given by Sornay did the reforestation efforts supported by the
(1937). Establishment of the Fonds Forestier F.F.N. But even with the reduction of the area
National (FFN) resulted in the availability of large annually planted to conifers, the proportion of
funds for re- Douglas-fir in coniferous plantations established
since 1980 remained nearly
5° 0°
5° constant, with an average of about 30% (De
Champs 1997b), until the end of the millenium.
The first Douglas-fir plantations financed by
50° 50° the FFN were made with seed purchased from
FRANCE American firms, but practically without control of
origin. These imports of seed of uncontrolled
origin con- tinued until 1966, the year the EEC
Nor th issued directives aimed at controlling the origin of
Atlantic Ocean
imported seed. Establishment of 90 ha of
Douglas-fir seed orchards between 1974 and 1990
will provide an important source of seed of known
45° 45°
origin.
Bay of Biscay Douglas-fir finds its best development in
Douglas-fir regions with a warm Atlantic climate as in the
1,068,150 m3
(79% of total volume) Bretagne, and at low and medium elevations up to
831,150 m3
(61% of total volume)
Mediter r anean 800 m in the west and southwest of the Massif
Sea 0 100 km
Central. But it also
5° 0° 5°
17. Letter from P. Bouvarel, Centre de Recherches INRA de
Figure 3.3 Standing volume (m3) of Douglas-fir in France (from De Nancy, dated May 5, 1970.
Champs 1997a). 18. Letter from J.-Ch. Bastien, INRA Centre des Recherches
d’Orléans of 7 August 1992.
Chapter 3. Areas of Introduction 49
19. Ibid.
performs well in other parts of the country, as for
example in the Cevennes and the Beaujolais. The
species is now planted throughout France except
for the Mediterranean region in the south, and the
Massif Landais in the southwest. Only the variety
menziesii is considered suitable for planting in
France. In the words of J.-Ch. Bastien,19 “The
reason for the success of Douglas-fir in France
are: its rapid and sustained growth, its high
production (13 m3/ ha/yr), its plasticity, the
absence of parasites or seri- ous pests, and the
mechanical qualities of its wood little changed by
fast growth. Its only fault is its
great attractivness to roe deer.”

Switzerland
In 1861, the Swiss Forestry Association formed
the “Kommission für Anbauversuche mit
exotischen Holzarten” (commission for trials with
exotic tree species) (Charbon 1991). The first
plantings of Douglas-fir in Switzerland were
perhaps made short- ly after the commission was
established. In any event, its seed inventory of
1865 lists Douglas-fir under the name Abies
Douglasii Lindl. (Schwager 1979). Based on the
age of the oldest Douglas-fir in the country, Hans
Burger (undated, cited by Schwager 1979, p.
91) presumed that the species was introduced to
Switzerland between 1860 and 1870.
The oldest forest plantings of Douglas-fir on
record were made on privately owned land in the
period 1874-1876 near Küssnacht. They were
mixed plantations of Douglas-fir, eastern white
pine and Norway spruce. A pure stand of
Douglas-fir (0.53 ha) was established by the same
owner in the spring of 1888 with 2-2 seedlings
(Coaz 1897). They had been grown from seed of
unknown provenance purchased in Erfurt,
Germany (Krutina 1927). That stand, situ- ated at
630 m elevation, grew so well that Dr. Coaz of
Bern made special reference to it in a 1905 speech
before the German Dendrological Society. Two
small plantations of Douglas-fir were established
in the City Forest of Biel in 1886 and 1893 with
2−2 and 2−1 stock, respectively. In 1926, Badoux
reported on the development of these three stands,
and pointed out that their volume production was
superior to that of Norway spruce on comparable
sites. The two
Begin 1992) have shown that Douglas-fir
oldest Douglas-fir plantations in the French performs well overall on the best sites and also on
part of Switzerland were established 1886 poorer sites. Diez and Bürgi believe that
near Lausanne (Charbon 1991). Interim provenances from elevations 500 to 700 m in the
reports on their develop- ment already
appeared at the beginning of the 20th century
(Curchod 1901, Buchet 1913).
Trees in all these plantations belong to the
va- riety menziesii but their provenance is
unknown. In an effort to learn more about the
seed origin of early plantations, Berney
(1972) used regressions of relative DNA
content of embryo cells on latitude to analyse
samples from a stand at Boesingen near Biel.
On the basis of that analysis, he concluded
that the Boesingen stand came from parents
situated between lat 44° N and 47° N in the
coast ranges of Oregon and Washington.
Although Douglas-fir appeared to hold
much promise for use in Swiss silviculture,
planting of the species was greatly reduced
from about 1930 to the early 1970s because of
concerns about the presence of Rhabdocline
pseudotsugae and Phaeocryptopus gaeu- mannii,
first noticed in 1914 and 1925, respectively. In
spite of that setback, coastal Douglas-fir has
become the most important introduced tree
species in the country. Pseudotsuga
represented 43% of exotic tree species in
Swiss forests in 1986. Its role is modest,
however, because estimates indicate that
Douglas- fir occupies less than 0.1% of
Switzerland’s forested area. Notwithstanding
its marginal importance, Douglas-fir has been
increasingly planted by Swiss foresters since
the 1970s. In 1986, nearly half of all Douglas-
fir plantations were in the 1- to 20-year age
class, and most of the other half were in the
60+ age classes (Bürgi and Diez 1986).
Douglas-fir plays an important silvicultural role
by the conversion of coppice forests on the
south side of the Alps, which is a consequence
of its relatively rapid juvenile growth, its high
volume increment, and the fact that Douglas-
fir can be planted success- fully on drier sites
for which a dearth of suitable indigenous
species exists (Buffi 1987).
Few Swiss studies have been concerned
with the productivity of Douglas-fir and its
relation to site characteristics. However,
recent investigations (Diez and Bürgi 1991,
50 Douglas-fir: The Genus Pseudotsuga
Washington Cascades are probably best suited for of the Forest Administration Ort near Gmunden,
Switzerland. The trials at Copera in the Ticino recordkeeping of the plots was discontinued and
show general superiority by Washington over
Oregon provenances (Buffi 1987).

Austria
The oldest Douglas-firs on record in Austria are
in the City Forest of Bregenz, Vorarlberg. They
were planted in 1876 in an abandoned nursery
during a training session for forest guards
(Rannert 1959a). Soon thereafter, the Austrian
Forest Experiment Station, largely through the
initiative of Adolf Cieslar, began to undertake
systematic trials with for- eign tree species,
including Douglas-fir. Cieslar ob- tained Douglas-
fir seed directly from G. B. Sudworth of the
Forestry Department, U.S. Department of
Agriculture (Rannert 1979), and later from
com- mercial sources. The origin of the seed is
unknown, except for the fact that shipments
included seed belonging to both the coastal and
interior varieties. Cieslar subscribed to the idea
that establishing many small plantations
distributed throughout the country would provide
more useful information than a few large
plantations because of the diversity of the Austrian
landscape. The first plantations, dating back to
1886, were concentrated in the Alps and the north-
western part of Austria. Cieslar (1898) attributed
the surprisingly satisfactory growth on his trial
plots to high amounts of precipitation, high
relative humid- ity, and good soils. Douglas-fir
performed well even in higher elevations of the
Austrian Limestone Alps (Oesterreichische
Kalkalpen), although the rate of growth was
slower than at low altitudes. At the turn of the
century 89 plots existed throughout the former
Austrian-Hungarian Monarchy encompass- ing
elevations from 120 m near the Adriatic Sea to
over 1,600 m in the Tyrolian Alps (Cieslar
1901). The report by Zederbauer (1919) on the
Austrian trials with exotic trees shows that in
1916, a total of 142,497 Douglas-firs had been
planted on 134 plots. The end of World War I
resulted in the dissolu- tion of the Austro-
Hungarian Empire. Thus, only 43 Douglas-fir
plots remained in the Republic of Austria
(Rannert 1979). But with one notable excep- tion,
a one-hectare stand in the Reindlmühl District
many were subsequently lost. Planting of since 1851. Trees planted in 1851 to 1863 had
Douglas-fir almost ceased until the 1950s, been import- ed from German nurseries (Holm
although Cieslar (1920) had argued that enough 1940). Introduction of the species into Danish
evidence existed to conclude that the climatic forests began in 1866. The area stocked with
conditions of Austria would permit Douglas-fir amounted to
satisfactory growth of the species. Another
advocate of Douglas-fir was Schwarz
(1932a,b) who, based on a study of site
conditions within the range of coastal
Douglas-fir, distinguished three climatic
growth regions for the cultivation of Douglas-
fir in Austria: the Vorarlberg region, Ober-
und Niederösterreich region, and the
Burgenland-Steiermark region. He
emphasized that only the coastal variety
should be considered for planting in Austria.
Renewed interest in Douglas-fir began with
the attempt by the Austrian Forest Experiment
Station in 1956 to salvage what was left of the
plots with foreign tree species established by
Cieslar. In a sur- vey of exotic trees that
covered all of Austria over a period of 14 years
(Rannert 1958, 1959a, 1959b, 1960; Minelli
1967; Rannert 1973, 1979) the data showed
that Douglas-fir occupied 108 ha in 266
localities. The largest concentration of
Douglas-fir (52 ha) was in Niederösterreich.
The majority of the trees (37%) were in the
age class 1–20, and those in the age classes
61–80 and 81–100 amounted to only 6%.
Information provided by the survey
demonstrated that coastal Douglas-fir shows
excellent growth when planted on suitable
sites. That was demonstrated by long- term
growth trials with coastal Douglas-fir with
unknown provenances in eastern and north-
central Austria. The mean annual increment at
age 70 was
18.5 m3/ha (Kristöfel 2003). The interior variety, on
the other hand, performed poorly and is
unlikely to be planted in the future. An
extensive program of provenance trials (Günzl
1987) that began at the end of the 1970s was
designed to provide a basis for a much wider
use of Douglas-fir in the forests of Austria
than in the past.

Northern Europe
Denmark
Douglas-fir has been planted in Danish parks
Chapter 3. Areas of Introduction 51
silvicultural practices
3.7 ha in 1882. That area grew to 10.1 ha by 1892,
btoy2179.012h, ato 72.2 ha by 1912, and to 154.5
ha by 1922 (Fabricius 1926). Planting in the next
25 years increased the area of Douglas-fir to
1,550 ha. Included in this figure are probably not
more than 10 ha stocked with interior Douglas-fir
(Thulin 1949). Madsen20 estimated that the area
occupied by the species has remained about at the
1947 level. That estimate appears to be
corroborated by an estimate of 20,000 m3 for the
annual cut of Douglas-fir in Denmark (Moltesen
1988).
Origin of the oldest Douglas-firs in Denmark is
unknown. Trees for the 1866 plantings were
obtained from Scotland and probably represented
progeny of trees grown from the seed shipped in
1826 by David Douglas. Holm (1940) cited
excerpts from a 1911 letter by the Danish seed
dealer Johannes Rafn, in which he expresses his
belief that the early Danish Douglas-fir
plantations were established with seeds from
stands in California and Oregon. Rafn based that
belief on the fact that early seed shipments came
from the firm Sonntag & Co. in San Francisco.
But that assumption is open to question. Some of
the seed supplied by Sonntag & Co. may also
have come from Washington. Rafn himself made
contact in 1902 with a seed dealer in Washington
from whom he purchased 100 kg seed in 1902,
150 kg in 1903, 250
kg in 1905, 500 kg in 1906, and 800 kg in 1909.
Thus,
many of the Douglas-fir plantations established in
the early part of the 20th century in Denmark
most likely represent Washington provenances.
Performance of Douglas-fir in the early planta-
tions was so satisfactory that Oppermann wrote in
1922, “About the year 1900 cultivation of the
green Douglas-fir was as yet in its experimental
stage. Since then the species has attained a secure
position in our forestry and the reason for
enlarging its cultivation are strong.” That view
was challenged by the ap- pearance of
Phaeocryptopus gaeumannii in Douglas-fir stands
throughout the country (Buchwald 1940). Larsen
(1940), however, considered pessimistic pre-
dictions about the future of Douglas-fir in
Denmark to be unjustified. He contended that
selecting of unsuitable sites and improper
Norway: before 1928, 34,030; 1928-1939,
were probably the primary causes for the 130,080; 1940-1949, 76,180; 1950-59, 20,000; 1960-69,
spread of the disease. Later developments 260,000. Heiberg (1978) examined 102 plantations in
proved him to be correct. Although Douglas- 1976 and 1978, ranging in age from 12 to 80
fir has found serious competitors in Sitka years, throughout Norway south of lat 63° N
spruce and grand fir, it seems to have the edge (Figure 3.4). Most of the plantations were small
in the driest parts of Denmark, the Jutland and of unknown provenance. Many were in steep
heath region, with annual precipitation of 500- and remote country
550 mm (Henrickson 1956, Oksbjerg 1965).
0° 8° 16° 32°
Barents
Norway Sea
0100 200 km
Single Douglas-firs have been planted in
western Norway since about 1870 (Nedkvitne
68°
1964). Heiberg (1978) listed location, height,
and diameter of 27 solitaires planted from
1879 to 1925. The tallest, planted in 1883,
SWEDEN
was 40 m high. The tree with the largest dbh,
118 cm, had been planted about 1880.
Plantations have been established since the 64°

turn of the century, but data on the area


stocked with Douglas-fir are not available. FINLAND
Børtnes (1970) gave the following figures on NORWAY
numbers of Douglas-fir planted in western
60 °
Oslo

20. Letter dated 4 November 1993 from S.F. Madsen, Danish ESTONIA
Figure 3.4 Locations of Douglas-fir
Douglas-firstands
stand in Norway (from
Forest North Sea
and Landscape Research Institute, Lynby Heiberg 1978).
52 Douglas-fir: The Genus Pseudotsuga
and had not received any subsequent 20°

silvicultural treatment.
FINLAND
Successful plantations appear to belong mainly
to provenances from the northern end of the
coastal variety’s range. The interior variety is
growing too slowly in Norway to be of value as a
forest tree (Hagem 1931). The same conclusion
was reached by Robak (1967) after more than 30
years of nurs- ery studies in western Norway. He SWEDEN
recommend- ed against the use of provenances
from interior British Columbia because of their
slow growth and great susceptibility to
Rhabdocline pseudotsugae and Phaeocryptopus
gaeumannii after outplanting in the field.
Moreover, in nurseries where frost heaving is a
problem, the small size of seedlings from interior
British Columbia provenances makes them par- 60°

ticularly prone to damage. Twenty-year results of NORWAY


a trial by the Norwegian Forest Research Institute Stockholm
with provenances from British Columbia showed
best growth for those from coastal sources, such
as Babine Range, and poorest performance for
Göteborg
those from interior sources, such as Prince George
(Heiberg 1978).
In general, Douglas-fir has performed best in
the Malmö
southern coastal region and the western coastal re-
gion northward to about lat 61° N. Heiberg Douglas-fir Stands
believed failures have been more common than
successes with Douglas-fir, and that the
principal cause of
failure was absence of suitable strains of mycor- inventory of Douglas-fir in Sweden by Lemoine and
rhizae, rather than frost damage or diseases. Both Wirten
Phaeocryptopus gaeumannii and Rhabdocline pseu-
dotsugae, however, have caused severe damage to
plantations (Nedkvitne 1964). A statement made
by Heiberg in 1975 is probably still valid: “Even
though Douglas-fir seems to be a promising tree
in the successful plantations, it has been
capricious in our trials, and its introduction is still
in the trial and error stage. Several more years and
plantations are needed before we may be able to
decide if it has a future as a commercial tree at
these latitudes.”

Sweden
The first Douglas-firs were planted in Sweden
about the middle of the 19th century. The first ac-
tual stand of Douglas-fir was established in 1880
on the Rössjöholms estate (Refn 1965). An
Figure 3.5 Douglas-fir stands in Sweden, totaling 100 ha
(from La Moine and Wirten 1988).

(1988) shows that small stands of the


species had been established throughout the
next 100 years. The plantations range from
Jämtland in the north to Skåne in the south
(Figure 3.5). The northernmost plantations,
Muråsen and Avardo, are at Frostviken, lat
64°30’ N. The inventory lists 96 stands
covering
99.85 ha. The total area occupied by the
species, however, may be somewhat larger.
Lemoine and Wirten cautioned that their
inventory should not be considered as a
complete catalogue because ad- ditional
stands are likely to be discovered on private
forest lands.
Lemoine and Wirten established permanent
sam- ple plots in 13 of the 96 stands and
control plots in nearby Norway spruce stands.
Stands ranged in age from 10 to 60 years
when plots were established.
Chapter 3. Areas of Introduction 53

After remeasurement of the plots in 1992, Douglas-fir stand


Eriksson and Widerlund (1992) concluded that the
NORWAY MT Mustiler SB Solböle
RK Ruotsynkylä AU Aulanko
results did not permit a generally valid PH Punkaharja
comparison between vol- ume production of KV Kivalo VP Vilppula VJ Vesijako EV Evo
LP Lapinjä rvi
Douglas-fir and Norway spruce.
Some of the Douglas-fir plantations have shown 68°
68°
re- markable growth with a total production of
slightly more than 650 m3/ha of stemwood in 60
SWEDEN
years.
Sweden’s climate severely limits choice of
prov- enances suitable for that country. 66° KV
66°
Experience has shown that provenances from the
major part of the range of the variety menziesii RUSSIA
will not survive, or they grow very poorly. Some
of the old stands that seem to be well adapted to
growing conditions in central and southern 64°
64°
Sweden originate from south- ern British
Columbia seed sources, or are progeny from FINLAND
Douglas-fir grown in Denmark (Martinsson
1990). Performance of these old stands and initial
62° VP
results from recent provenance trials (Martinsson PH 62°
and Kollenmark 1993) suggest that provenances VJ
AUEV
from the northern part of the range of Douglas-fir LPMT
RK
are those that offer the most promise for 050 100 km

SB050100 mi
successful cultivation in Sweden.
60°
60°
Finland
20° 30°
The oldest plantation of Douglas-fir in Finland
was established in 1905 in the Mustila Arboretum.
The seed source of trees in that plantation is
indicated
as Quesnel, British Columbia. Subsequently, trial clearly maladapted to the semi-maritime climate of
plantations of Douglas-fir were established at Finland and tend to become stagnant at ages of more
Solböle (1924, 1926, 1927, 1937). Ruotsynkylä than 30 years (Tigerstedt 1990).
(1924, 1927,
1942), Aulanko (1927) and Punkaharja (1926,
1927, 1938). The locations of these trial plantations
between lat 60° N and 62° N are several degrees
above the northern limit of the natural range of the
species (Figure 3.6).
All surviving trees in 1980 belonged to prov-
enances from interior British Columbia and
Alberta, except for one from Washington (Lähde
et al. 1984). Early growth of Douglas-fir appears
to have been rather poor. In 1956, Heikinheimo
published a list of Douglas-fir stands in Finland,
and wrote in a later publication (Heikinheimo
1957) that growth of these stands was not very
promising. Most of these interior provenances are
Figure 3.6 Douglas-fir stands in Finland (from Lähde et al. 1984).
the second northernmost of the surviving
The survey of Finnish Douglas-fir stands provenances was already considered to be the best
by Lähde et al. (1984) indicated that volume (Cajander 1926, Ilvessalo 1926). Mean volume
produc- tion by age 55 years may approach production of the 5 most productive stands
that of the native Pinus sylvestris and Picea ranging in age from 50 to 55 years was 400 m3/ha.
abies on good sites, but the quality of In Finland, Douglas-fir suffers from frost and
Douglas-fir is poorer. Crooked, bowed and Rhabdocline pseudotsugae. Lähde et al. (1984)
leaning stems are common in all Douglas-fir con- cluded Douglas-fir “is a species which has
stands. At age 75, the stand at Mustila had received much interest but recent measurements
780 m3/ha, with a dominant height of 28 m, indicate that Douglas-fir cannot be recommended
and a dominant dbh of 38 cm. At an early age, for widescale commercial use in Finland.”
54 Douglas-fir: The Genus Pseudotsuga

Mediterranean Europe species represents only about 0.1% of the


country’s total forest cover. Fontes et al. (2003)
Portugal
investigated the environmental factors affecting
Douglas-fir
A single Douglas-fir planted on the slope of the were used, nota- bly Douglas-fir (Luis 1989). As a
Castelo dos Mouros, another next to the Fonte dos result of that pro- gram, Douglas-fir occupied about
Passarinhos, and a third tree of a P. menziesii var. 7,000 ha in 1989. Forecasts were that the area planted
glauca, mark the introduction of the species into to Douglas-fir would have increased as a result of the
Portugal (Gomes and Raposo 1939). The oldest of “Programa de Acçáo Forestal” (forest action Program)
these trees, the one near the Castelo dos Mouros, to more than 15,000 ha by 1997 (Rego and Alvares
was planted sometime between 1844 and 1849. 1988).
As to the origin of the Castelo dos Mouros and The fast growth of Douglas-fir, its ability to grow
Fonte dos Passarinhos trees, Gomes and Raposo under a wide variety of conditions, and to regenerate
suggested that they were progeny of trees raised naturally, has shown its great potential for planting in
in Scotland from the seed shipped by David the mountains of central and northern Portugal (Goes
Douglas. Other early introductions are Douglas- 1991, Louro and Cabrita 1989, Luis 1989, Diniz 1969).
firs planted 1871 in the Forest of Buçaco (Peres Although that potential is recognized, the
1964), and several trees dating to about 1876 in
the Pedras Salgades Park (Coutinho 1936,
Gaussen 1944).
The first forest plantations of Douglas-fir were
es- tablished by the Forest Service in the Serra da
Estrella in 1904 (Freitas 1989), and in the Serra do
Gerês in 1906 (Coutinho 1936). But in general, little
use was made of Douglas-fir until the last quarter of
the 20th century, except for individual trees planted
in parks and gardens, and small plantations
established by the Portuguese Forest Service
(Carvalho 1965).
In 1976, the area of Douglas-fir plantations
was less than 300 ha (Goes 1991). A few large
planta- tions were established in the late 1970s
and early 1980s at Bornes and Malcata, increasing
the area of Douglas-fir to 4,200 ha.
Through the “Fundo de Fomento Florestal
(Forest Development Fund) a large afforestation
program was initiated at the beginning of the
1970s. For that program various exotic conifers
productivity and, based on these factors,
modeled Douglas-fir productivity to provide
information for future afforestation projects.
They estimated that an area of 250,000 ha
exists where Douglas-fir trees could be
planted and will exceed 17 m dominant height
at age 30 years. This would correspond to 8%
of the existing Portuguese forest area. The
best sites for Douglas-fir growth are those in
the north coastal to central regions at altitudes
between 500 m and 1,000 m, and with a
moisture deficit (precipitation minus
evapotranspiration) above 1,000 mm. Areas
with acceptable sites for Douglas-fir growth
are in the north and center of Portugal at 700
to 1,000 m elevation, and with a moisture
deficit above 400 mm.

Spain
Single Douglas-firs planted in the Province of
Vizcaya in the Basque country shortly before
and after the turn of the century probably
represent the first intro- ductions of the
species to Spain. Pellon (1962) refers to a 74-
year-old Douglas-fir with a height of 45 m
and a dbh of 72 cm, and in 1966, to another
solitary Douglas-fir tree that had attained a
height of more than 40 m at age 60. He noted,
however, that forest plantations dating to the
turn of the century do not exist in Vizcaya.
The oldest plantations of Douglas-fir
appear to be in the province of Gerona in
northeastern Catalonia on three private
ownerships. One of these planta- tions, on the
Serrat estate, dates to 1926. Plantations on the
Mas-Joan Garolera estates were established in
the 1950s. Arenas (1962), who gave a detailed
ac- count of the Douglas-fir plantations on
these estates, estimated that yields of 500 to
600 m3/ha may be expected with rotations of
45 to 60 years on sites of medium fertility.
Overcutting and forest fires have led to the
disap- pearance of much of the native forest,
degradation of sites, and development of
scrublands in the moun- tains of northern
Spain (Arenas 1962). Douglas-fir is
considered to be a promising species for
rehabilitat- ing such sites, as well as for
abandoned agricultural lands at sites above
600 m that are unsuitable for the economically
more profitable plantations of
Chapter 3. Areas of Introduction 55

Monterey pine and Eucalyptus globulus. Estimates Allegri’s (1962) perhaps somewhat exuberant
of the extent of land available for planting statement, “We can claim that Douglas-fir has
Douglas- fir range from 300,000 ha (García, 21 now victoriously found its way into Italian
Vega 1990) to 550,000 ha.22 silviculture,”24 apparently reflected the growing
The early trials resulted in far less satisfactory interest for the spe- cies stimulated by the results
results with the interior than the coastal variety of of Pavari’s trials. The view, however, that
Douglas-fir (Arenas 1962). Provenance trials with Douglas-fir had a place in Italian silviculture was
coastal Douglas-fir initiated in 1978 are intended not universally shared. Giacobbe argued in a
to identify seed sources best suited for use in series of papers (1942, 1963, 1967) that the tree’s
Spain (Vega 1990). liability to damage from climatic agents and
About 5,000 ha of Douglas-fir plantations ex- disease, notably Rhabdocline pseudotsugae and
isted in Spain in 1993 distributed through Phaeocryptopus gaeumannii, would preclude it from
Catalonia, Rioja, Navarra, the Basque country, general cultivation in Italy. Merendi (1956, 1965)
and Galicia.23 By contrast, De Champs (1997), disputed the case against planting of Douglas-fir
based on informa- tion received in 1993 from G. and stated that the risk of disease had been
Vega of the Centro de Investigaciones Forestales exaggerated. The view in favor of Douglas-fir has
at Louridan Pontevedra, puts the acreage for apparently prevailed because planting it has
Douglas-fir in Spain far higher than 5,000 ha. continued. The tree has proved to be
According to Vega, the area covered by Douglas- particularly useful for two purposes. One is
fir is estimated to be 30,000 ha of which 21,000 afforestation of abandoned ag- ricultural land in
ha are in public and 9,000 ha in private owner- mountainous regions to prevent erosion and to
ship. The annual area planted to Douglas-fir in the keep the land productive. The other is the
1980s was about 800 ha. Lack of an adequate conversion of forests of low production to higher
supply of seed has been an impediment to more production (Morandini 1968). Increased
extensive planting of the species (Vega 1990). production is made possible by the growth rate of
Douglas-fir, which enables it to outproduce both
Italy
native conifers and broadleaves (Susmel 1962,
A single Douglas-fir planted 1858 in Tuscany in Ciancio et al. 1980). For instance, Cristofolini
the parks of Moncioni southwest of Montevarchi, (1968) reported mean an- nual increment of 11.7
and one or two at Brolio northeast of Siena m3/ha, and current annual increment of 17.3
represent the earliest known introduction of the m3/ha for 22-year-old Douglas-fir in Liguria.
species to Italy (Bernetti 1987). The tree at Morandini (1961) noted that on good sites in the
Moncioni was 16 m high and had a dbh of 33 cm Apennines, Douglas-fir can attain volumes of
in 1883 at age 25 (Booth 1907). In 1918, at age 60 500 m3/ha at age 30.
that tree was 26 m high with a dbh of 78 cm The Castanetum and the Fagetum are phytocli-
(Hickel 1922). The first actual plantations were matic zones best suited for planting Douglas-fir
established between 1885 and 1890 at with optima in the cool subzone of the
Vallombrosa, followed by others at Masseto and Castanetum and the warm subzone of the Fagetum
Bivigliano near Florence (Ciancio et al. 1980). (Pavari 1958). Upper altitudinal limits for
In general, however, foresters showed little in- cultivating Douglas-fir are given by Susmel
terest in Douglas-fir until the extensive work with (1962), as follows: 500 m in the eastern Alps, 800
the species by Aldo Pavari, Professor of m in the central and western Alps, 1,200 m in the
Silviculture central Apennines, and 1,400 m in the southern
Apennines.
at Florence University. He established 98 experi-
mental plantations of Douglas-fir in the Alps; the central and northern Italy is suitable for the species
northern, central, and southern Apennines; and the (Pavari and de Philippis 1941, Pavari 1958).
Mediterranean region between 1922 and 1938.
These trials demonstrated that a variety of sites in
21. Letter from José García Salmerón, Instituto Forestal de
Investigaciones y Experiencias, Madrid, dated 15
September 1970.
22. Letter from Dr. Ramon Elena, Instituto Forestal de
Investigaciones y Experiencias, Madrid, dated 27 October 1993.
23. Ibid.
24. Translated from the original: “Possiamo affermare che la
Douglasia e ormai entrata vittoriosamente nella selvicoltura
italiana” (Allegri 1962).
56 Douglas-fir: The Genus Pseudotsuga
Only the coastal variety of Douglas-fir is and other unspecified causes. Use of provenances
suitable for cultivation in Italy. Initial experience suitable for Croatia (Pintarić 1967) may help to
indicates that Oregon and northern California obtain higher yields from Douglas-fir.
provenanc- es hold the most promise, although
Slovenia
some south- ern Washington provenances also
perform well (Morandini 1968). The largest The earliest plantings of Douglas-fir, in about
concentration of plan- tations is in the northern 1890, were made when Slovenia was still part of
Apennines, especially in Tuscany, with about the Austro-Hungarian Empire (Miklavzic 1951).
4,000 ha (Bernetti 1987). Most of the stands were These plantings, however, were either solitary
established between 1950 and 1970. More trees or very small plantations. In 1926, Douglas-
recently, reforestation activities declined in fir occupied 16 hectares. Nearly all of these were
general. Douglas-fir is planted only on a limited in the Maribor district with the exception of 0.3
scale, mainly to replace degraded silver fir (Abies ha in the Ljubljana district (Urbas 1926). More
alba) stands in the Apennines.25 Douglas-fir has been planted since but mostly in
According to statistics of the Italian Forest mixture with Norway spruce, European silver
Service, 429 stands of Douglas-fir occupied 1,315 fir and eastern white pine (Cokl 1965). Statistics
ha in 1963. By 1980, the area planted to Douglas- on area currently occupied by the species in
fir was estimat- ed by Ciancio et al. (1980) to be Slovenia are unavailable.
somewhat in excess of 10,000 ha. The Italian Greece
National Forest Inventory of 1985 lumped
The 1919 planting of Douglas-fir in the arboretum
Douglas-fir together with other exotic conifers.
of Vitina in the Peloponnesus marks its
Thus only estimates are available, which place the
introduction to Greece. Since 1960, some trial
area stocked with Douglas-fir still at slightly over
plantations have been established with the species.
10,000 ha.
A current annual increment of 15 m3/ha between
Croatia ages 16 and 20 was recorded at Pertouli in central
Douglas-fir is important in Croatia for increases Greece and of 17 m3/ ha at Chalkidike in eastern
in timber yield through amelioration of degraded Greece. But, aside from these trial plantations
forests (Pavle 1967). That applies particularly to Douglas-fir was planted only on an extremely
rehabilitation of degraded sites in the Karst region small scale and covers probably not more than
(Piškorić 1960). 100 ha (De Champs 1997b).
Planting of Douglas-fir must have already Turkey
taken place in about 1890 in Croatia because The introduction of Douglas-fir to Turkey dates to
Klepac (1962) mentions a 70-year-old stand with the years after World War II. The favorable results
a standing volume of 678 m3/ha in the River of a few small plantations near the Black Sea
Valley of the Gorski Kotar region. Apparently, coast suggested that Douglas-fir might be a
Croatia has more Douglas-fir stands than any of valuable tree species for the country’s
the other republics of the former Yugoslavia, but reforestation program in that region (Simsek
actual figures for the area occupied by the species 1977). A comprehensive prov- enance trial was
in Croatia are unavailable. Klepac (1962) pointed initiated in 1971 to identify potential seed sources
out that productivity in Croatian Douglas-fir suitable for Turkey.
stands is less than that given in British (Hummel
and Christie 1953) and German (Schober 1955) Cyprus
yield tables. He attributed the lower produc- tivity A small plantation of coastal Douglas-fir has been
of Croatian Douglas-fir stands to unfortunate established in a sheltered stream bed at about 762
choice of planting sites, poor silvicultural m altitude on this island in the eastern
practices, Mediterranean Sea. The trees were healthy, but
their growth had been slow (Streets 1962).
25. S. Nocentini, University of Florence, letter of January 17, 1994.
Chapter 3. Areas of Introduction 57

Eastern Europe questionnaires. The next inventory covered only


stands older than 10 years by 1953 in both the
Czech Republic Czech and Slovak parts of the country (Hofman
and Heger
The history of the introduction of Douglas-fir up greater than shown by returned
to 1919 in the Czech Republic, made up of
Bohemia and Moravia, has been described in
detail by Nozicka (1963). Interest in Douglas-fir
began as early as 1828 when F.G. Rietsch of
Zbraslav near Prague discussed the tree in the
journal “Forst- und Jagdneuigkeiten.” The oldest
Douglas-fir on record in the Czech Republic was
planted in 1843, as a 2- or 3-year-old seedling in
the American Garden of the Chudenice
Arboretum in western Bohemia near Klatovy.
That seedling was part of a shipment of exotic
trees from the Booth Nursery in Flottbeck,
Germany. Another record of an early planting is
from Gross-Skall (Hruba Scala) Estate near
Wartenberg. Here, two 5-year-old Douglas-firs
were planted in its Bukovina Park. The two trees
came from the Booth Sons Nursery in Flottbeck
(Anger 1879). Hofman (1964) lists several
locations where Douglas-fir pre- sumably had
been planted in the period 1850-1870, but details
are lacking.
Although the earliest plantings of Douglas-fir
consisted only of solitary trees, small plantations
of the species in forests began to be made in the
late 1860s. An inventory of Douglas-fir stands
published by Hofman and Heger in 1958
indicated establish- ment of 36 stands from 1868
to 1878, and of 95 stands between 1878 and 1888.
Planting of the species at- tained a first peak in
the decade 1908-1918 and a second peak in the
years 1928-1938. This peak was followed by a
decline in planting of the tree until 1954. That
year marked the beginning of renewed planting
activities with Douglas-fir (Hofman 1962).
The first attempt at an inventory of Douglas-
fir in what was then Czechoslovakia was made
by Polansky (1934) by means of questionnaires
sent to forest districts throughout the country. He
recorded 24 ha of pure stands and 78 ha of
stands where Douglas-fir was mixed with other
species. Polansky doubted that responses to the
questionnaires pro- vided accurate figures,
however, and suspected that the actual area
stocked with Douglas-fir was two or three times
1958). It showed 447 stands in the Czech
Republic, with an area of 120.5 ha. But even
that inventory apparently failed to include all
existing stands of Douglas-fir. Sika (1979), in a
later inventory restricted to the Czech Republic,
found that the area occupied by Douglas-fir
older than 60 years (i.e., for planta- tions
established before World War I), amounted to
245 ha. The total area stocked with Douglas-
fir in 1976 was 1,990 ha. On nearly half of
that area stood plantations in the 1- to 10-year
age-class. Stands are small, ranging in size
from 0.07 ha to about 2 ha with an average
size of 0.15 ha. In 1982, plantations less than
10 years old occupied 1,125 ha, those in the
11- to 60- year age-class 774 ha, and stands in
the 60+ age-classes 621 ha (Sika 1983).
More recent data on forest tree species in the
Czech Republic26 show an increase from
2,520 ha stocked with Douglas-fir to 3,800 ha
at the end of 1991. Distribution by age classes
had also shifted, with 990 ha in the 10-year
class, 2,252 ha in the 11- to 60-year classes,
and 558 ha in the 61- to 130-year age-classes.
The largest concentration of Douglas-fir is in
southwestern Bohemia (Figure 3.7). Standing
volume of Douglas-fir in the Czech Republic
in 1991 was nearly 370,000 m3.
The first Douglas-firs planted in Bohemian
forests were raised from seed purchased from
C. Geyer, a forester turned seed dealer in
Karlshafen on the Weser River in Germany.
Afterwards, seeds were directly imported
from the USA (Hofman 1962). Information on
the amount and origin of seed pur- chased
before 1910 is not available. Between that
year and 1948, 961 kg of seed were imported
and 120 kg were obtained from Douglas-fir
stands in Czechoslovakia. A large amount of
seed (4,635 kg) was imported in the years
1955-1957, at the begin- ning of a 10-year
plan aimed at establishing 20,000 ha of
Douglas-fir plantations, equivalent to 0.5% of
former Czechoslovakia’s forested land.
Hofman (1962) pointed out that the 1,081
kg of imported and domestic seed should have
been suf- ficient to establish at least 1,000 ha
of plantations.
58 Douglas-fir: The Genus Pseudotsuga

14° 16°

Douglas-fir (ha)
Over 200
150-199
100-149
GERMANY POLAND 50-99

CZECH REPUBLIC

50° Prague 50°

GERMANY Brno

03060 km

AUSTRIA SLOVAKIA
60 mi
030

14°
16° 18°
Figure 3.7 Area stocked with Douglas-fir in Czech Republic (from Vancura 1993).

But, according to him, only 196 ha of stands older ticularly suited for use in Czech forests (Hofman et
than 10 years existed in 1953 in Czechoslovakia. al. 1964, Sika 1982). Hofman (1962) noted that the
Likewise, the fulfillment of the 1956 ten-year plan growth of Douglas-fir in the Czech Republic and
fell far short of its goal of 20,000 ha of Douglas- Slovakia is comparable to that of the species in other
fir plantations by 1966, at least judging by the countries of central and western Europe, an indica-
results of the 1976 inventory. The reasons for this tion that very favorable conditions exist for
are poor quality seed, unsuitable provenances, increased planting of the species in both countries.
poor choice of planting sites, lack of proper Apparently, efforts were made to plant more
silvicultural treatment, and insufficient protection Douglas-fir, with the ultimate goal of having
against wildlife damage (Hofman 1962, Sika 120,000 ha in the Czech Republic—that is 5% of all
1979). its forested land—stocked with Douglas-fir (Sika
Studies of the growth of Douglas-fir in 1981).
Bohemia date back to the last quarter of the 19th
century, and the beginning of the 20th century. Slovakia
The Imperial Austrian Forest Research Institute Douglas-fir was introduced to Slovakia27 as an
Mariabrunn estab- lished experimental plots in or- namental tree about 1830, when the country
Bohemia, then still a part of the Austro-Hungarian was still part of the Austro-Hungarian Empire.
Empire. These were estab- lished in 1883 (Hofman Use of the species as a forest tree began toward
1962), 1902, and 1905 (Žabka 1946). These and the end of the 19th century, when professors at
later investigations by the Czech Forest Research the Mining Academy in Banka Stiavnica
Institute at Zbraslav indicated that Douglas-fir can established trial plots in forest stands. These early
be grown successfully, particularly plantings were made with
in the rolling hill country of the Czech Republic (Sika
1979). Early plantations included both the interior provenances par-
and the coastal varieties of Douglas-fir
(Hejtmanek 1952), but only the latter performed
satisfactorily. Trials are under way to identify
26. Copy received February 8, 1993 from K. Vancura,
Czech Forest Research Institute Zbraslav.
27. The account of Douglas-fir in Slovakia is based on the
letter of Sept. 1, 1993, from Dr. Peter Tavoda, Forest
Research Institute, Zvolen, Slovakia.
Chapter 3. Areas of Introduction 59

trees of unknown
provenanc- es, some of 18° 20° 22°
UKRAINE
which were un- suitable,
Haromhuta
and plantings failed soon SLOVAKIA
after establishment. Other 48° AUSTRIA
provenances used in the
early trials adapted well to
the new environment and
grew vigorously. The
oldest existing stands in HUNGARY
1993 were 100 years old. Zalaegerszeg
In 1993, Douglas-fir cov-
ROMANIA
ered about 1,200 ha (0.18%
Iharosbereny
of Slovakia’s forest lands)
Douglas-fir stands
and the number of pure 46° CROATIA
0100 km
SERBIA
stands was small. Usually, 02040 mi

Douglas- fir is mixed with 18° 20° 22°


Norway spruce, European
silver fir,
or European larch, and Figure 3.8 Location of oldest Douglas-fir stands in Hungary (from Harkai 1975).
some-
times with Scots pine and
grand fir. Most of these stands are younger than of present-day Hungary, 28 near Haromhuta in
50 years, and the proportion of Douglas-fir gener- northeastern Hungary, and in Iharosbereny and
ally ranges from 10% to 30%. The species has
been planted throughout Slovakia at altitudes
from 200 to 1,000 m, but the best growing
conditions are found at elevations between 400
and 600 m. Stands on the best sites have reached
heights of 43 to 45 m.
In general, experience with Douglas-fir in
Slovakia has been favorable and it may be expected
that its use will be continued in Slovac forests on a
modest scale.
Hungary
In 1877, Pausinger suggested considering the
intro- duction of Douglas-fir as a means of
increasing the yield in some Hungarian forests.
Shortly afterwards, Bedö (1878) recommended
trials with Douglas-fir. He appeared to have been
influenced by Booth’s (1877) book, in which that
author extolled the merits of Douglas-fir. The
introduction of the species into Hungary began
with the procurement of several kilograms of seed
in the years 1880-1882 by the Hungarian National
Forestry Association (Harkai 1975). Guidelines
for trials with Douglas-fir were drawn up shortly
afterwards (Dietz 1885, Marosi 1885).
The oldest Douglas-fir stands in the forests
Zalaegerszeg in western Hungary (Figure 3.8)
stem from around the turn of the 19th century.
Their provenance is unknown. Interest in the
species ap- peared to have waned in the first
half of the 20th century, but increased after
World War II because of the government
policy to solve the softwood shortage through
conifer plantations. That discus- sion was
reflected by the establishment of numerous
trial plots of Douglas-fir until 1970 (Harkai
1975). Nonetheless, the area stocked with
Douglas-fir has remained small. In 1961,
Papp published a list of Douglas-fir stands in
Hungary, most of which were less than one
hectare. The total area occupied by the species
was given as 37 ha. That area has increased
only moderately since 1961. It amounted to
353 ha in 1990.29
That Douglas-fir occupies only such a small area
is surprising in view of the fact that the
species finds favorable growing conditions on
sites with brown forest soils 30 where
precipitation is above 600 mm (Papp 1961,
Szöny 1963, Harkai 1971). These are sites

28. Hungary lost nearly 90% of its forest lands to other


countries after World War I. As a result, the initial
plantations of Douglas-fir established by the Hungarian
Forest Service are now outside the territory of present-day
Hungary.
29. Letter of Sept. 13, 1993 by Dr. Mátyás Csaba, Univ. of Sopron,
Hungary.
30. Corresponds approximately to mollisols.
60 Douglas-fir: The Genus Pseudotsuga
31. Letter of Sept. 13, 1993 by Dr. Csaba Mátyás, University of
where the yield of Douglas-fir is considerably Sopron, Hungary.
higher than that of native conifers (Bano 1963).
Inventories have shown 60-year-old stands of
coastal Douglas-fir with standing volumes ranging
from 800 to 1,000 m3/ ha (Szönyi and Nagy
1968). In a comparative trial at the Budafa
Arboretum in southwestern Hungary, Douglas-fir
reached a mean annual increment of 22 m3/ha at
age 20, which is twice the m.a.i. of tradition- ally
planted Scots pine (Gergacz and Csaba 1993).
Under field conditions, 250-350 m 3/ha of standing
volume has been measured at age 40 in the same
region depending upon initial spacing (Harkai
1987). Although the yield potential and the range
of sites suitable for Douglas-fir are sufficiently
well known, the species has not been planted
more widely for several reasons. Parts of the
country have a combina- tion of semiarid climate
and soils that result in site conditions where native
conifers and broadleaves show much better
growth than Douglas-fir (Majer 1980, Gergacz
and Csaba 1993). Even on sites favor- able to the
growth of Douglas-fir, the rate of initial mortality
in plantations is high because of the fre- quency
of spring droughts. In general, more than
50% of seedlings in a plantation are replanted. 31
Fencing of plantations is necessary for at
least 20-25 years. Damage by deer through
stripping the bark with their antlers threatens
the existence of unprotected stands. In the
inadequately protected comparative test of
conifers at Zalaerdod, Douglas-fir proved to be
the most susceptible to game damage and, thus,
the least productive (Harkai 1981).

Serbia
The share of conifers amounted to slightly more
than 7% in the forests of Serbia in 1962. Plans were
made to increase that share and to include Douglas-
fir with the species considered for planting (Marić
1962).
Douglas-fir represented a miniscule part of
Serbian coniferous woodlands. Petrović listed six
small stands of Douglas-fir known to him. One of
those stands in the Avala State Forest near
Belgrade was established in 1911 when the
Serbian Forest Service converted oak coppice to
conifers (Marković 1951). The performance at
Avala State Forest prompt-
ed Radulovi (1960) to advocate wider use of Washington. The remainder stemmed from British
the species. Development of Douglas-fir Columbia. Data on the amounts of seed imported
plantations established at that time in central since 1970 were not given. Petkova’s review
and eastern Serbia were described in 1973 by merely indicates that several provenances from
Stamenković and Miscević. Information on western
provenance of Serbian Douglas-fir stands is
probably unknown because it is not men-
tioned by those authors. However, one of the
initial stands is of such quality that it was
selected as a seed stand (Marić 1962).
Information about the current status of
Douglas-fir in Serbia is unavailable.

Bosnia and Herzegovina


Professor Konrad Pintarić of the faculty of
Forestry, University of Sarajevo, initiated an
experiment in 1963 with provenances from the
Pacific Northwest. Seedlings were outplanted
in 1966 at Batalovo, about 20 km west of
Sarajevo. In 1997, the standing volume
projected to per hectare, ranged from 125 to
235 m3 (Ballian et al. 2003). Additional trials
were estab- lished at Crna Lokva (44° 51′ N,
16°51′ E), elevation 665 m (Pintarić 1989,
Godevar et al. 2003); Gostovic (44°23′ N
18°08′ E) elevation 411 m; and Blinje (43°50′ N,
18°03′ E), elevation 951 m. The performance
of Douglas-fir in these limited trials suggests
good po- tential for introduction to Bosnia and
Herzegovina (see additional information in
chapter 4).

Bulgaria
A review by Petkova (2004) of the history and
per- formance of Douglas-fir in Bulgaria
indicates that it has become well adapted to
site conditions in that country. The French
forester Felix Wogeli established the first
Douglas-fir plantation in 1906 with seed of
unknown origin at 700 m elevation in the
Stara Planina mountain range. Other early
plantings in parks in Sofia, Koprivschtitza,
and the Rila monas- tery region constituted
mixtures of Douglas-fir with hardwood and
other softwood species.
Planting of Douglas-fir ceased during the
period 1930–1956, but recommenced after the
beginning of renewed seed imports in 1956.
Bulgaria imported 12,949 kg of Douglas-fir
seed between 1956 and 1969. Most of the seed
originated from sources in western
Chapter 3. Areas of Introduction 61

Washington and a single prov- 24° 26° 28°

enance from Oregon (Forest UKRAINE


Trial plantations Very favorable Favorable sites
Grove) were used for 48°
plantings in the period 1970– Suceava
HUNGARY
1985.
In addition to imported Oradea
MOLDOVA
seed, domestic seed has
already been used. Cone
crops of the oldest Douglas- ROMANIA
46°
fir stand in the Kazanluk
Forest District have yielded
sufficient amounts of seed
permitting the establish-
ment of second-generation
plantations.
Bucharest
Douglas-fir has shown SERBIA
44°
good growth, especially at

l
BULGARIA
eleva- tions between 800 and
22° 24° 26° 28°
1,200 m in the Stara Planina
Range. Inventories of 43
plantations in this mountain
range indicated
that 20-year-old Douglas-fir Figure 3.9 Douglas-fir stands in Romania. Dots denote location of stands: dark areas = sites very
trees on moist sites with favorable for planting; lighter marked areas = sites favorable for planting.
north- ern exposures had a
mean an- nual increment
between 15 and
18 m3/ha. But m.a.i. at that age was much lower, 6 Planting of Douglas-fir continued on a very limited
3
to 8 m /ha, on dry, warm sites. scale until about 1950.
Petkova stated that the total area of Douglas-
fir plantations amounted to 12,664 ha in 1985, but
gave no source for that figure. Inventory data of
the Bulgarian Forest Service cited by her
indicated that Douglas-fir was found on 6,792 ha
in 2000. She did not explain the large decrease in
the area occupied by the species, however. Large
increases in the population of roe deer in the
second half of the 20th century, and drought
conditions in the de- cade 1983-1993, may have
resulted in considerable losses of trees, but these
events are unlikely the sole reasons for a
reduction of nearly 6,000 ha stocked with
Douglas-fir over a relatively short span of time.

Romania
Douglas-fir was introduced into Romania in the
period 1870-1880, as plants or seeds of unknown
provenance. All that is known about that early in-
troduction is the fact that the plant material came
from Austria (Lazarescu and Ionescu 1964).
From 1959 to 1962, Ionescu and Lazarescu
(1966) inventoried and studied stands
throughout Romania that contained Douglas-
fir. The coastal variety of the species was
represented in 41 stands covering a total of 59
ha, the interior variety in 8 stands on about 7
ha. Nearly all stands were mixed, with the
share of Douglas-fir varying between 10%
and 90%. Coastal Douglas-fir was represented
in all age-classes from 15 to 70 years. Based
on the findings from their study, Ionescu and
Lazarescu identified regions suit- able for
cultivating the coastal variety in Romania
(Figure 3.9). The best growing conditions
exist in the southwestern part of the country at
elevations between 300 and 1,000 m.
By chance, many of the early plantings
were made with provenances of the coastal
variety that is well suited for growing
conditions in Romania. After the encouraging
performance of coastal Douglas- fir in these
early plantings, 23,382 ha of plantations were
established between 1960 and 1970 with seed
imported from a dealer in the Pacific
Northwest (Violeta Enescu 1979).
Unfortunately, many of these plantations
failed, and in 1993, only about 13,000 ha
stocked with Douglas-fir were left. The
imported
62 Douglas-fir: The Genus Pseudotsuga
seed was presumably of unsuitable provenance Schwappach from British Columbia, Washington,
and, thus, became a major contributing factor to and Great Britain.
the planting failures.32 The first paper on the introduction of Douglas-
To reduce the chance of such failures in the fu- fir to Poland appeared before the turn of the
ture, provenance trials have been initiated (Violeta century (Tyniecki 1891), and as early as 1912,
Enescu 1984). However, the need for imports of Sokolowski reported that yield of coastal
seed is greatly reduced because seed can be Douglas-fir was su- perior to native conifers of
harvested in sufficient amounts from existing the same age on similar sites. In 1926, Suchocki
Douglas-fir stands of good quality in Romania. remeasured plots of coastal Douglas-fir
Additionally, seed can be obtained from 37 ha of established by Schwappach in three for- est
seed orchards, and rais- ing planting stock through districts in western Poland. The results indicated
vegetative propagation seems possible.33 that volume production of these stands was equal
The Romanian Institute of Forest Tree to that of British Douglas-fir stands of the same
Improvement recommends planting coastal age. Suchocki noted that performance of the
Douglas- fir on sites that have at least 750 mm of interior va- riety was unsatisfactory, and
annual precipitation and a mean annual recommended against its planting. Bieler (1935)
temperature of 7–9°C, and avoiding sites with reported a total yield of 626 m 3/ha for a 54-year-
stagnant air or ex- cessive drought in the spring. old stand planted in 1881 in the Poznan region, a
The interior variety of Douglas-fir has grown yield that according to him is produced there by
satisfactorily in parts of Romania where a European silver fir at age 62, and by Norway
continental climate prevents cul- tivation of the spruce only at age 72.
coastal variety (Haralamb 1971). But any future Maciejowski (1950, 1951) concluded, after a com-
use of the interior variety is likely to be on a prehensive review of experience with Douglas-fir
miniscule scale. in European forests, that its potential for use in
Currently many silviculturists are oriented to- Polish forests had been underestimated. He
ward the use of indigenous species even if they believed that Douglas-fir could play a role
are less productive than Douglas-fir. Researchers rehabilitating the coun- try’s forests ravaged in
will need to demonstrate that Douglas-fir is a World War II. However, he made a strong plea
remark- able species that should have a place in that Douglas-fir should not be planted in
Romanian silviculture. Bialowieska and parts of Bialystok and
Augustowo Forests to preserve the unique
Poland
character of these forests.
Douglas-fir was first introduced to Poland in 1833 Stands of exotics, as well as files of pertinent
by Count Stanislaw Wodzicki, who had it planted information, had been destroyed in many
as an ornamental on his estate in Niedwiec near instances during World War II. In 1959, Bialobok
Cracow.34 The oldest stands of the species were published an inventory of foreign tree species
established in three waves, 1879–1880, 1891–1895, made in the post- war years in an effort to
and 1907–1910. determine and salvage those that had remained.
Most of these stands are in territories formerly He mapped and described 103 stands of Douglas-
under Prussian and Austrian administration in the fir covering 191 ha. He pointed out, however, that
western part of the country and in south-central his list was probably incomplete. A few years
Poland, respectively (Bialobok 1959). Seed later, an update of the original exotics inventory
origins of the stands are unknown. Maciejowski was published (Bialobok and Chylarecki 1965). It
(1950) presumed that they were established with showed the existence of 1,169 stands of Douglas-
seed obtained by fir (Figure 3.10) comprising a total area of 1,405
ha.
32. Valeru Enescu, Romanian Institute of Forest Tree Improvement, An excellent account of Douglas-fir in Poland
Bucharest, letter of 23 February 1993.
is provided by Chylarecki’s (1976) study of 84
33. Ibid.
34. L. Meynartowicz, Polish Academy of Science, Institute stands selected from 1,136 sites throughout
of Dendrology, Kórnik, letter of 15 June 1993.
Poland. His data show that Douglas-fir grows
best in the coastal region along the Baltic Sea
with its oceanic
Chapter 3. Areas of Introduction 63

climate and annual precipita-


tion between 600 and 700
RUSSIA LITHUANIA
mm. Other regions where Gdansk

Douglas-fir grows well are the 54°


54°
lower eleva- tions in the
Sudeten Mountains, Silesian
Beskide Mountains, and Szczecin
Bialystok
Carpathian Mountains.
Volume yield surpasses that of
almost all the indigenous Warsaw BELARUS
Poznan
conifers. At age 80, mean 52° 52°

stand volume is 640 m3/ha, Lodz


which corresponds to the
volume of the best Norway
spruce stands and exceeds GERMANY POLAND
Lublin
yield obtainable from Scots Wroclaw

pine stands by nearly 40%.


The maximal yield of
3 Katowice
Douglas-fir of 850 m /ha is
close to the pro-
Rzeszow
ductivity of the best stands of 50° CZECH REPUBLIC
50°
Krakow

European silver fir. An UKRAINE


Douglas-fir stands
updated and enlarged version
of the 1976 study was SLOVAKIA

published in German
(Chylarecki 2005). 15° 20°

Figure 3.10 Douglas-fir stands in Poland (from Bialobok and Chylarecki 1965).
Tumilowicz (1967)
recom- mended that coastal
Douglas-
fir should even be grown in the western and future use of Douglas-fir in Poland. L. Mejnartowicz
central Masurian Lake region of northeastern (letter of July 10, 1993) indicates that in 1993 about
Poland. Only in the eastern part of the region are 1,000 ha were stocked with Douglas-fir, and that
temperatures in winter too low for survival of opposition by environmentalists is unlikely to prevent
the coastal variety. Tumilowicz based his future use of the species in Polish forests.
conclusions on an inventory of Douglas-fir
stands older than 30 years in 24 forest districts in
the Masurian Lake region. Stands estab- lished
in the years 1884-1897 had standing volumes that
would fall between site classes III and IV of the
McArdle and Meyer yield tables.
Dominik’s (1963) statement, “Douglas-fir is
little known by our foresters while lovers of our
nature protection are ardent opponents of it,” is
probably not applicable anymore, at least in
regard to the first part of the statement. Polish
foresters have become more knowledgeable about
the species, and Polish participation in the IUFRO
international Douglas-fir provenance trial
(Mejnartowicz 1976) is a clear sign of interest for
Baltic States
Lithuania
A map from Jankauskas (1951) shows 27
locations throughout Lithuania where stands
of Douglas-fir are situated (Figure 3.11). The
oldest of these stands were established in
1900 to 1910. Most of the existing plantations
appear to belong to the interior variety, which
is considered to be completely acclimatized in
Lithuania (Navasajtis 1966).

Latvia
Douglas-fir has been planted in Latvian parks
since the middle of the 19th century. The first
forest plan- tations were established between
1900 and 1902 (Pirags 1990). In 1968, 10
stands of Douglas-fir of unspecified size
existed in the forest districts Talsi, Rezekne,
Ogre, Dobele, and Dangavpils, as well as
numerous groups and solitary trees in other
dis- tricts and parks (Pirags 1968). The area
occupied by Douglas-fir older than 20 years
was very small, but new plantations have been
established since 1965 (Pirags 1979).
64 Douglas-fir: The Genus Pseudotsuga
about 25 ha. Trees on 5 of
LATVIA the 25 ha stemmed from
BELARUS
Mažeikiai domes- tic seed collections.
Birzai
56°
Kretinga
56° Domestic seed crops are
fairly good. Douglas-fir
S

Klaipeda begins to produce seed at


LITHUANIA about age 20,, and con-
tinues to produce cone
crops at 2- to 3-year
intervals (Laas 1967).
Etverk (1978) attrib- uted
Kaliningrad Oblast RUSSIA Kaunas the small area covered by
Trakai District
(former East Prussia) Douglas-fir to a shortage of
Vilnius
Aukštadvaris seed and lack of system-
Alytus atic provenance
Douglas-fir stands
experiments. The oldest
54° POLAND 54°
03060 km BELARUS stand in 1978 was about 70
03060 mi
years old. Nearly all stands
22° 24° 26°
belonged to the inte- rior
variety, referred to as ei-
ther variety glauca or
variety caesia. The interior
variety is
Figure 3.11 Douglas-fir in Lithuania (from Jankauskas 1951). fir of un- specified size stood 15 forest districts and in
several arboreta and parks (Margus 1961, 1963). By
Both the interior and coastal variety have been 1978, the area occupied by the species had increased
planted in Latvia, the latter being the more to
produc- tive one under Latvian growing
conditions. Douglas- fir is outperforming both
Norway spruce and Scots pine, with a m.a.i. of 11
m3/ha at age 60 (Pirags 1979).
Estonia
The earliest attempt to grow Douglas-fir in
Estonia, situated between lat 57°30’ and 59°40’ N,
long 21°45’ and 28°15’E, was made in 1880 in the
environs of Tallinn. That initial attempt was
unsuccessful, prob- ably because of unsuitable
seed (Margus 1961). First efforts to find suitable
seed sources were made by Count Berg, a private
forest owner, and Mr. M. Sievers, chairman of the
Baltic Forestry Association (Baltischer
Forstverein). In spite, of the encouraging results
of a few small plantations established early in the
20th century, the species received little attention
until 1954. That year, H. Taimre kindled renewed
interest in Douglas-fir through a paper about a
stand of Pseudotsuga in Kaarepere Forest District.
In 1958, small stands covered about 6 ha in
seven forest districts. In 1963, groups of Douglas-
considered to be suitable for
all of Estonia, but the
coastal variety is only suitable for the western
part of the country (Laas 1967).

Former Soviet Union


Belarus
Following a survey of the performance of
conifers introduced into Belarus, Ivanova
(1963) recom- mended Pseudotsuga menziesii
var. glauca for use in forest plantations.
Skutko (1966) evaluated the performance of
several provenances of the interior variety,
which he referred to as varieties caesia and
glauca, in plantations where they grew mixed
with Siberian larch, Norway spruce, and scots
pine. He noted that trees of variety glauca
were inferior to them and thus should not be
considered for use in forest plantations.

Ukraine
Both the coastal and the interior variety, as
well as some cultivars of Douglas-fir, are
grown in Ukrainian gardens and parks (Lipa
1940), but only the variety menziesii is used in
forest plantations. The excel- lent growth of
Douglas-fir in the western Ukraine prompted
Brodovich (1964) to recommend collecting seed
from stands of the species in the Ukraine to
Chapter 3. Areas of Introduction 65

Table 3.5 Growth and productivity data of Douglas-fir in the Ukraine (from Brodovich 1967).

Mean Volume increment (cm3)

Age (y) Number Basal Form Standing Mean Current


Height (m) Diameter (cm) of area 2 factors volume 3
stems/ha (cm ) (0.001) (cm /ha) (annual) (annual)
10 4,6 4,7 641
20 14,1 14,3 2136 34,3 492 238 11,9 27,0
30 22,8 22,5 1258 50,0 478 545 18,2 32,0
40 29,2 29,4 862 58,5 475 810 20,2 24,8
50 34,1 34,8 664 63,2 473 1018 20,3 18,6
60 37,8 38,8 555 65,6 473 1167 19,5 12,8

establish more Douglas-fir plantations. He wrote in 1976, sequently in the Soviet Union was reviewed by
“Almost 70-year-long trials in acclimatization and Kalutskii et al. (1981). The Central Research
propagation of Douglas-fir in the Soviet Union proved that Institute for Forest Genetics, Breeding, and
the green Douglas-fir35 in the western Ukraine is one of the Selection, estab- lished at Voronezh in 1971 paid
most rapidly growing, valuable and promising forestry particular attention to the introduction of Douglas-
species.” Data on growth and productivity of Douglas-fir in fir and other fast- growing exotic conifers in the
the Ukraine reported by Brodovich (1967) are reproduced in western regions of the former USSR. A map
Table 3.5. prepared by that institute shows a division of the
Douglas-fir has been grown in plantations in the country into 6 zones accord- ing to suitability for
Carpathian region of the Ukraine since 1906 and, in 1985, growth of introduced species. Zones are numbered
covered 1,555 ha (Rudenko and Derzhanovskaja 1985). The consecutively 1 to 6 from north to south. Planting
best sites for Douglas-fir in the western Ukraine are north, of Douglas-fir is recommended for zones 3, 4, and
northeast, and northwest slopes. Natural regeneration is 5. According to Tkatchenko (1958), W. Sievers
abundant under favorable site conditions ranging from had suggested the line Leningrad – Moscow
20,000 seedlings per hectare (Brodovich 1978) to 70,000 to – Saratov as the eastern boundary for cultivating
100,000 plants per hectare (Rudenko and Derzhanovskaja Douglas-fir in Russia.
(1985)). According to Brodovich (1978), Douglas-fir has Plantations of Douglas-fir have been
completely adapted to conditions in the western Ukraine and established in European Russia since the last
is far more productive than the native Norway spruce. decade of the 19th century. Reports on Douglas-
fir in the forest-steppe region (Vehov and Vehov
Russia
1962, Akimockin 1964, Lukin 1966, Biryukov
As elsewhere in Europe, arboreta were the first 1971) indicate that the spe- cies shows good
places where Douglas-fir was planted. The oldest growth. In the south-central part of Russia, growth
Douglas- firs in the Nikitsky Botanical Garden in and productivity of Douglas-fir exceed those of
the Crimea are thought to have been raised from most native and introduced species. Dudetskaja
seed brought between 1830 and 1840 from the and Lukin (1977) emphasized that the species has
former Russian colony at Fort Ross in California survived the severest winters there.
(Wulff 1926). In 1914, the Moscow Botanical The Black Sea coastal zone of the Caucasus
Garden contained sev- eral Douglas-firs planted and the Crimea are important resort areas where
about 1880 (Meyer 1914). The history of the exotic conifers have been introduced primarily for
introduction of Douglas-fir and other exotic aes- thetic reasons (Dusha 1977). But the species
conifers in tsarist Russia, and sub- also has great potential for timber production in
35. “Green Douglas-fir” = Pseudotsuga menziesii var. menziesii.
that region (Sud’ev 1980), as is apparent from the
performance of a plantation of Douglas-fir about 5
km from the Black Sea coast. At age 32, it had a
standing volume of 602 m3/ha (Dudarev et al.
1975).
66 Douglas-fir: The Genus Pseudotsuga

Asia between 1940 and 1974, reaching a total of


47,126 ha (Table 3.6). Most of that area, 90.8%
India was in State forests. In 1964, Rotorua
Conservancy, with 26,188
A trial in the Punjab on a hot, bare, and rocky ha, contained 55.6% of the country’s Douglas-fir
hillside at 2,134 m elevation with exotic species resource. At that time, somewhat more than two-
included 100 Douglas-firs. By 1951, only three thirds of New Zealand’s Douglas-fir stands were
trees had sur- vived. Das and Chand (1958) on the North Island, most of them concentrated in
attributed the failure of Douglas-fir to the the Rotorua Conservancy. In 1974, there were
mismatch between species and site. about 27,000 ha of Douglas-fir plantations less
than 20 years old. Of these plantations, those on
Sri Lanka more than 11,000 ha (almost all in State forests)
Douglas-fir was planted in the 1920s in Sri Lanka had been estab- lished between 1966 and 1970.
(formerly Ceylon) at elevations between 1,219 The peak of planting Douglas-fir in that 5-year
and 1,829 m. Troup, in 1932, described the results period was followed by a marked decline in the
as “unpromising.” following years (Fraser 1978). It is unlikely that
these 1974 figures reflect the extent of previous
Southern Hemisphere planting because some planta- tions failed. For
The areas of introduction of Douglas-fir in the instance, of the more than 18,000 ha of Douglas-
south- ern hemisphere are New Zealand, fir planted between 1920 and 1936 in Kaingaroa
Australia, Chile, Argentina, and eastern and Forest, a third of these plantations failed
southern Africa. Except for Africa, the areas of (Kirkland 1969). Another example comes from
introduction in the southern hemisphere all have the Westland Conservancy. In the 1950s,
parts where the species has been grown Douglas-fir was planted in State forests at an
successfully. annual rate of 100 to 200 ha, and that was
increased to 600 to 700 ha in the 1960s (Allan
Southwestern Pacific
1978). However, only 461 ha of Douglas-fir were
New Zealand and Australia are the regions of ear- recorded for the State forests of
liest introduction of Douglas-fir to the southern the Westland Conservancy in 1974.
hemisphere. Aside from France, Germany, and The New Zealand Ministry of Forests annually
the United Kingdom, New Zealand has by far the publishes, “A National Exotic Forest Description”
larg- est acreage of that species outside its natural (Neumann 1993), which includes data on stocked
range. In Australia, planting of Douglas-fir has areas and age-class distributions for Douglas-fir
remained limited. Lack of suitable sites and for all 78 districts in the country. The data for
climatic limitations are probably the principal 1990 and 1991 are contained in Table 3.7. These
reasons that the tree has not found wider use in data cover only
Australian forestry.
Table 3.6 Number of hectares stocked with Douglas-fir in New
New Zealand
Zealand in 1974 (From FraserPost
Up to 1940 1978).
1940 Mixed-age Total
Douglas-fir was introduced to New Zealand in 1859 (ha) (ha) classes (ha) (ha)
by J. B. Acland of Canterbury (Anonymous 1994). From North Island
about 1870, the species was planted as an ornamental and Auckland 17 2,055 – 2.072
shade tree on farms in the South Island (Streets 1962). Rotorua 13,974 12,214 – 26,188
Planting of Douglas-fir in State forests began in 1897, when Wellington 429 5,677 – 6,106
it was among the species used in the first trials of exotic South Island
trees on the Kaingaroa Plains in central North Island Nelson 771 4,883 – 5,654
Westland 7 454 – 461
(Kirkland 1969). The area planted to Douglas-fir was small
Canterbury 457 2,610 26 3,093
until 1917: 450 ha (Spurr 1961). By 1940, the area of
Southland 764 2,788 3,552
16,419 30,681 26 47,126
Douglas- fir plantations had increased by almost
16,000 ha. Nearly double that area was planted to
Douglas-fir
Chapter 3. Areas of Introduction 67

Table 3.7 Number of hectares stocked with Douglas-fir in New Zealand, 1940 to 1991.
19401 19741 19902 19912
Auckland 3
Northern4
17 2,072 5 2
Rotorua Central 13,974 26,188 36,309 31,571
Wellington Southern 429 6,106 1,614 1,587
North Island 14,420 34,366 37,928 33,160
87,8% 72,9% 57,2% 55,9%
Nelson/Marlborough 771 5,654 11,583 11,439
Westland 7 461 556 556
Canterbury 457 3,093 5,292 4,663
Otago/Southland 764 3,552 10,936 9,473
South Island 1,999 12,760 28,367 26,131
12,2% 27,1% 42,8% 44,1%
New Zealand Totals 16,419 47,126 66,295 59,291
1.Fraser 1978.
2.Neumann 1993.
3.As listed by Fraser 1978.
4.As listed by Neumann
1993.

Table 3.8 Douglas-fir age class distribution in New Zealand as of April 1, 1991 (from Neumann 1993).

Age class (y) 1–10 11–20 21–30 31–40 41–50 51–60 61–80
Hectares 15,661 13,047 16,909 5,673 859 2,944 4,198
Percent 26.4 22.0 28.5 9.6 1.4 5.0 7.1

about 90% of the total area occupied by Douglas-


species are 900 m in the North Island and 750 m
fir, however, because the forest survey does not
in the South Island (Kirkland 1971). Wind
include holdings of less than 100 hectares. If these
exposure is a major limitation to survival and
areas are included, the total area stocked with
satisfactory growth at all altitudes. Because New
Douglas-fir for all of New Zealand was 71,066 ha
Zealand is a relatively narrow country oriented at
in 1990, and 65,478 ha in 1991.36
right angles to the prevailing west to
The data in Table 3.7 show a substantial
southwesterly air flows, many localities
increase in the area stocked with Douglas-fir from
experience, strong, desiccating or salt-carrying
1974 to 1990, particularly in the central North
winds. Lack of sufficient precipitation is not a
Island re- gion, and in Nelson/Marlborough,
limiting factor, except for parts of Canterbury and
Canterbury, and Otaga/Southland. They also
Southland (Prior et al. 1963, Revell 1978).
demonstrate that the area occupied by Douglas-fir
Douglas-fir has been grown in New Zealand
decreased, the result of increased harvesting to
since the second half of the 19th century, but
supply the booming log export industry.37 A
records of seed origin were not kept until 1927.
breakdown of Douglas-fir stands by age (Table
The 6,650 ha of Douglas-fir planted from 1915 to
3.8) shows that most stands (76.9%) were less
1928 in the north- ern part of Kaingaroa Forest
than 30 years old.
are presumably from western Washington seed
Douglas-fir can be grown successfully in many
sources. The New Zealand Forest Service
parts of the country, provided that appropriate
imported 3,223 kg of seed from 1927 to 1974
sites are chosen. In general, altitudinal limits for
(Table 3.9). During that same period 8,381 kg
the
were obtained from Douglas-fir stands in New
Zealand (Wilcox 1978; Table 3.10). Imports ceased
36. Dr. Colin O’Loughlin, Forest Consultant, Wadestown, in 1930; seed supplies from 1931 to 1964 were
Wellington, letter of April 21, 1993.
obtained entirely from New Zealand stands.
37. Ibid.
Major local seed sources have been Kangaroa,
Whaka, Golden Downs,
68 Douglas-fir: The Genus Pseudotsuga

Table 3.9 Imports of Douglas-fir seed to New Zealand since


the Stoke, Tapawera, Tadmor, and Motupiko dis-
1926 (from Wilcox 1978).
tricts. These shelterbelts in turn are thought to be
Year Origin Quantity (kg) progeny of the Mararewa Cemetery plantings
1927 Southern Washington 308 made about 1875 by John Stanley who brought
Salmon Arm, British Columbia 62 the seed from England. In general, the Douglas-fir
1928 Washington 27 in Golden Downs has a much narrower genetic
Washington 163
base than the stands in Kaingaroa (Wilcox 1978).
Washington 78
Further large im- ports of Washington seed were
1929 Coast Range, Oregon 91
made in 1965-66, and again in 1968 when J.
Coast Range, Washington 32
Spiers procured an assortment of seedlots from
Washington 322
Washington, Oregon, and California. Some coastal
Washington 54
Southwest Washington 272
Californian seed was imported in 1970 and 1972
1930 Cowlitz County, Washington 705
as a direct result of the observation that low-
West of Cascades, British Columbia 17
elevation Californian Douglas-fir was perform-
1956 Caspar, California 18 ing exceptionally well in provenance trials in New
1965 Pierce County, Washington, 150 m 227 Zealand. Experience has shown that provenances
1966 Pierce County, Washington, 150 m 204 of Douglas-fir from Oregon and California were
Humboldt County, California, 360 m 5 the best performers in New Zealand.38
1967 Meyer’s Flat, California 460 m 7 The future role of Douglas-fir in New Zealand
1968 Deadwood, Oregon 302 forestry is to some degree a matter of contention.
Bandon, Oregon 35 Of concern is the presence of Phaeocryptopus
Langlois, Oregon 39 gaeuman- nii, first detected in New Zealand in the
Langlois, Oregon 11 1960s. That fungus is thought to be largely
Snoqualmie, Washington 32 responsible for the severe decline of growth in
Snoqualmie, Washington 14 stands established be- fore 1940. But other
Snoqualmie, Washington 6 factors, such as insects, a series of dry summers,
Tahkenitch, Oregon, 240 m 6 unsuitable sites, and insufficient care of stands,
Mt. Rainier National Forest, Washington 97
may have been contributory causes (Groome
Pecwan, California, 240 m 11
1978, James and Bunn 1978).
Pecwan, California, 200 m 10
But economics is what weighs most heavily
1970 Swanton, California, 80 m 10
against the planting of Douglas-fir. Fenton stated
1972 Korbel, California, 80 m 13
in 1976 that in the context of national forest plan-
1974 Coquille River, Oregon, 200 m 45
Total
ning, the planting of Douglas-fir implies a delay
imports 3,223 in attaining production targets and a reduction in
profitability because the species takes
considerably longer than Monterey pine to reach
Tapawera, Coalgate and Queenstown. Kaingaroa
merchantable size (Fenton 1978). Because most
Forest alone was the source of 3,377 kg, an
of New Zealand’s State Forests have been
amount slightly higher than that of all imported
privatized, there is no overall policy within the
seed. From 1927 to 1930, 2,131 kg of seed were
forest sector about Douglas-fir.39 Past experience
imported, main- ly from Washington. Central and
suggests that the potential is greater for Douglas-
southern stands in Kaingaroa planted in 1930 to
fir than for Monterey pine at elevations above
1936, amounting to 5,600 ha, are from these
650 m in the North Island and above 500 m in the
Washington seed lots (Wilcox, 1978). Some of
South Island (Revell 1978). The great- est
the oldest stands in Golden
potential for expanding the Douglas-fir estate in
New Zealand exists in the hills and high country
of the South Island, where annual rainfall exceeds
800
Downs Forest originated from Washington seed.
Most of the Golden Downs stands, however, 38. Ibid.
origi- nated from seed collected from old 39. Ibid.
shelterbelts in
Chapter 3. Areas of Introduction 69

Table 3.10 Sources and quantities of Douglas-fir seed used in New Zealand 1926-1974 in kg (from Wilcox 1978).

Source 1926-34 1935-44 1945-54 1955-64 1965-74 Total


Imports 2,131 0 18 18 1,074 3,223
Kaingaroa 0 102 490 1,330 1,455 3,377
Whaka 166 110 121 37 11 445
Karioi 0 0 168 76 0 244
Golden Downs 11 10 122 293 543 979
Tapawera 197 122 162 186 138 805
Hanmer 0 2 271 154 149 576
Ashley 0 0 0 0 164 164
Coalgate 0 0 0 154 271 425
Palmside 0 0 0 37 37 74
Wanaka 0 0 0 35 54 89
Queenstown 0 0 22 32 359 413
Naseby 0 0 12 19 88 119
Tapanui 0 0 161 0 0 161
Dusky 0 0 39 69 81 189
Other sources 46 1 60 115 99 321
Total 2,551 347 1,628 2,555 4,523 11,604

mm. Ledgard and Belton (1985) and Belton Australia


(1991) have shown that Douglas-fir grows Statistics for Douglas-fir in Australia are not as
extremely well, at rates of between 20 and 32 complete as for New Zealand. Available data indi-
m3/ha/yr for fully stocked stands before age 50 in cate a total area of about 2,665 ha in State Forests
high country areas below 900 m elevation and (Griffin and Matheson 1978). Apparently few, if
with precipitation of more than 800 mm per year. any, Douglas-fir plantations are on privately
The eastern high country of the South Island has owned land. Nearly all Douglas-fir plantations are
about 200,000 ha of land physically suited to located in the southeastern part of Australia. The
forestry. Much of that land is partly or severely species has been planted on a trial basis in the
degraded under the present pastoral land use. In Mount Lofty Range of South Australia and in
any future large-scale afforestation projects in the Western Australia (Troup 1932, Streets 1962).
South Island high country, Douglas-fir would un- Douglas-fir was one of the secondary species
doubtedly be an important, if not the most in the planting program for the upper elevations of
important component of the forest scheme.40 the Southern Table Lands in New South Wales
The future prospects for New Zealand-grown (Streets 1962). According to Fenton (1967), 67 ha
Douglas-fir have received a boost because the were planted
declin- ing supply of Douglas-fir from the United to Douglas-fir between 1920 and 1960, and 437
States has raised the price of Douglas-fir wood in ha between 1961 and 1965. An anonymous article
international markets, thus improving the in the Australian Timber Journal (1965) cited by
economics of growing Douglas-fir in New Fenton (1967) reported plans for planting about
Zealand. A sign of confidence in the future of the 200 ha per annum in New South Wales. Margules
species in New Zealand was the fact that the New (1968) wrote that about 101 ha are planted
Zealand Forest Research Institute at Rotorua set annually to Douglas- fir on the Bago State Forest
up a Douglas-fir Research Cooperative in 1993 and that, in general, Douglas-fir is likely to
with the aim to foster continued research into the become one of the major species at altitudes of
silviculture, genetics, and use of Douglas-fir in 762 to 1,219 m in the Tumut and Bombala
New Zealand. districts. That forecast was unrealistic. Griffin and
Matheson (1978) gave the area stocked with
40. Ibid.
Douglas-fir as 1,350 ha, and added that New
70 Douglas-fir: The Genus Pseudotsuga
South Wales has no current planting program. In 65 ha in 1978 (Griffin and Matheson 1978). Insect
the Capital Territory, Douglas-fir grows pests of economic importance have not been
vigorously in the Cotter Valley at elevations observed
between 762 and 1,463
m. The species does not thrive, however, at lower
elevations in the vicinity of Canberra (Streets
1962). Douglas-fir was introduced to Victoria
about 1902 (Troup 1932), but was not planted on
State land until 1920. The Forests Commission
planted the tree on favorable sites in areas such as
Beechworth and the Air Valley (Margules 1968).
In 1962, Victoria had about 809 ha of Douglas-fir
plantations, of which 381 ha were in one block.
Most of the plantations were established between
1936 and 1942 at elevations between 457 and 914
m with seedlings of unknown provenance. Mean
annual precipitation for the plant- ing sites ranges
from 889 mm to 1,905 mm. Growth is best where
fog and low clouds are common and rainfall is
light but frequent. Shelter is essential for good
growth. Exposure to wind results in reduced
height growth but little blowdown (Streets 1962).
The Forests Commission of Victoria announced in
1957 that Douglas-fir would not be planted any-
more. The Forests Commission reversed that deci-
sion 5 years later when it indicated that the
species would be considered for suitable sites.
The planting of Douglas-fir reached a second
peak in the years 1965 to 1975. The emphasis
was on higher elevation sites where Monterey
pine suffers snow damage. Planting was again
suspended when the presence of Phaeocryptopus
gaeumannii was discovered. In 1978,
Douglas-fir occupied 1,250 ha (3,088 acres)
(Griffin
and Matheson 1978).
In Tasmania, Douglas-fir was probably intro-
duced during the gold rush of the 1850s. The tree
was planted as an ornamental until the Forest
Service in 1936 established Douglas-fir
plantations at Warrentina (Streets 1962). Annual
precipitation on planting sites ranges from 1,016
to 1,524 mm. Natural vegetation on these sites is
eucalypt forest dominated by Eucalyptus regnans
and E. obliqua. By 1956, the area planted to
Douglas-fir amounted to 94 ha (Fenton 1967).
Streets (1962) mentions that the extension of
Douglas-fir plantings was planned at the rate of 20
ha per year. These plans were not realized
because plantations of Douglas-fir occupied only
on Douglas-fir in Australia. In Victoria, Colorado. The interior variety was judged
wallabies have frequently damaged trees to unsuitable, and has not been used again (Streets
age 6 or 7 years (Streets 1962). 1962).
Douglas-fir in Australia belongs almost In all likelihood, Douglas-fir is going to remain
exclu- sively to the coastal variety of unknown an insignificant component of Australian forests.
provenance, except for trees in nine The prospects for future use of Douglas-fir have
provenance plantations established in been evaluated by Fenton (1967) as follows:
1972/1973. Spurr (1961) suggested that On end-use considerations only, there is little need for
provenances from the central Sierra Nevada Australia to plant Douglas-fir, as equivalent timber is
of California might be suitable for Australian potentially available from radiata pine and the object
conditions and should be tested at higher of expanded afforestation is to replace imports. To do
this by growing Douglas fir instead of radiata pine as
elevations in New South Wales and Victoria.
an exotic would take from 20 to 30 years longer, and
Preliminary results from 40 seedlots of the market preference for the species would have to be
IUFRO Douglas-fir collection under test in sustained by imports during this period to benefit
New South Wales, Victoria, and Tasmania, from its current market reputation. Apart from sites
however, indicated that provenances from which present limitations for radiata pine, there is no
compel- ling reason why afforestation of Douglas-fir
higher elevations in California are highly
should be extended in Australia.
susceptible to Phaeocryptopus gaeumannii.
Provenances from the coastal fog belt of Although the areas suitable for large-scale com-
northern California and southern Oregon mercial planting of Douglas-fir are limited, the
appear to be best suited for use in Australia. tree has use for shelterbelt and ornamental
The only report of the existence of a planting. According to Margules (1962) few
plantation with trees of the interior variety is species are more suitable for single-row
from Tasmania; shelterbelt planting than Douglas-fir because it
1.2 ha were planted at Stoodley Forests in has numerous small branches which, in the open,
1940 with are retained to ground level and carry vigorous
trees raised from seed imported from foliage to an advanced age.
Chapter 3. Areas of Introduction 71

South America region I, rotations of 60-68 years will be necessary to


cut trees with a dbh of 51 cm. In the plains, rotations
In contrast to the Southwest Pacific region, trial with the species was made by the Caja Agraria,
Douglas-
with the successful establishment of several Douglas-
fir had been planted on a relatively small scale in
fir plantations between 1939 and 1944 on its property
South America in the 1990s. That scale will near Loncoche (Weber and Gothe 1954). In spite of
probably change as the depletion of the encouraging results in these early trials, use of
indigenous forest resource progresses, and Douglas-fir as a plantation tree increased only slowly
reforestation efforts will be helped with until the 1960s. By 1967, Douglas-fir occupied be-
accelerated planting of fast-growing exotic tween 2,000 and 2,500 ha. The major share (1,500 ha)
species. or about 60% of the total area covered by Douglas- fir
Chile at that time in Chile, was concentrated in four estates
(Rocuant 1967). By 1977, the area covered by
Douglas-fir was introduced to Chile as an
Douglas-fir had increased to 5,449 ha (Diaz-Vaz and
ornamen- tal tree about 1895 (Rocuant 1967). A
Ojeda 1980), and to 9,000 ha by 1985 (Instituto
67-year-old Douglas-fir cut for stem analysis in
Forestal 1986).
1962 came from a park near Valdivia (Brun 1963).
Existing plantations are in a part of Chile that
The first planta- tion of Douglas-fir was
extends from lat 35° S in the north to about lat 42° S
established in 1914 by Hugo Weber on his estate,
in the south. Growth studies carried out in 1966/67 by
Bellavista, near Traiguen (Weber and Gothe
Rocuant (1967) and Contreras and Smith (1973)
1954). At age 50, trees in that plantation had
indicated impressive growth of Douglas-fir. Based on
reached heights of 35 m and diameters at breast
his findings, Rocuant recommended the area from lat
height of 63.5 cm (Rocuant 1967). Another early
35° S to lat 43° S for growing of Douglas-fir. But
because of differences in growth patterns of the would probably have to exceed 80 years because
species found in his 1966/67 study, he suggested a Douglas-fir grows more slowly there. In region II,
distinction between two regions: region I from lat a dimension of 51 cm is probably obtainable with
35° S to 38°30’ S, which includes the provinces
40- to 50-year rotations.
of
Maule and Malleco, and region II from lat 38°30’ Drawing on results from a series of experimen-
to tal plantations installed by the Instituto Forestal in
1962, and reports of the studies by Rocuant
(1967), Contreras and Smith (1973), and
Contreras and Peters (1982), the Instituto Forestal
produced a map of potential volume growth of
Douglas-fir according to homogenous edapho-
climatic units in the areas recommended for
planting of Douglas-fir (Figure 3.12). A note of
caution was added, however, by pointing out that
predictions of growth behavior under local
conditions will be difficult because of the extent
of the edapho-climatic units.
Provenance is undoubtedly another source of
variation in growth patterns. Seed being used in
Chile has come primarily from seed dealers in
the
Potential Growth (m3/ha/year)
15-20
10-15
5-10
36°
CHILE

37°

38°

39°

40°

41°

ARGENTINA 42°

43°
74° 75° 76° 77°
43° S, which encompasses the provinces of Cautin Figure 3.12 Potential volume growth of Douglas-fir in areas of
and Chiloe. In the coastal and Andean Cordilleras Chile recommended for planting the species (from Instituto
of Forestal 1986).
72 Douglas-fir: The Genus Pseudotsuga
United States. Carrasco (1954) mentions that (Fernandez 1964). A 13-year-old
planta- tions in the pre-Cordilleran zones of
Valdivia prov- ince were established with seed
harvested in the fog belt zone of southern
Washington and northern Oregon, supplied by
Manning Seed Company in Roy, Washington. In
general, information about the origin of seed
appears to be non-existent, and provenance tests
have been of limited scope. In a test of 14
provenances initiated by Rocuant in 1965, the best
results were obtained with provenances from the
coastal region of southwestern Oregon (Lopez
1973). Droppelmann (1986) analysed the
performance of 17-year-old Douglas-firs from 10
provenances in a trial near Valdivia. He found that
“near coast provenances” were superior to other
provenances in increment at the test site.
Planting of Douglas-fir is likely to increase in
the
future because it adapts well to the climate in
Chile, shows good growth, and produces wood
whose quality is considered equal or better than
that of Monterey pine (Weber 1953; Diaz-Vaz and
Ojeda 1980). Some of Douglas-fir’s silvical
characteristics contribute also to its growing
popularity. The species can withstand extended
periods of drought without major injury. Unlike
Monterey pine, it is much less likely to suffer
from snow breakage at sites where that is a
problem (Buch 1965). Douglas-fir has also been
shown to have potential for rehabilitating sites
degraded by certain agricultural practices (Buch
1978).

Argentina
In one of the first published accounts of Douglas-
fir in Argentina, Rodriguez (1960) mentions the
existence of several small plantations of the
species throughout the Andean-Patagonian forest
region, and of experimental plots in the National
Parks Nahuel Huapi (Isla Victoria) and Lanin.
Rodriguez stated that these plantations, roughly
situated be- tween lat 38° and 42° S and long 71°
and 72° W, were thriving and growing rapidly.
Information on the age of the plantations was not
given. Data on growth are available for young
stands only. A 20-year-old plantation near the
northern fringe of Lake Lacar about 20 km west
of San Martin de los Andes had a standing volume
of 634 m3/ha and a mean height of 18 m
plantation in the Cordilleran region in the a single plantation of Douglas-fir in subdivision H
northwest of the province of Chubut had a at 1,615 m altitude. Trees belonged to the interior
mean height of 15 m (Miglioli and Rozados variety of Douglas-fir. Their mean height and dbh
1972). at age 51 years were 9.75 m and 19.8 cm,
The performance of the plantation near respectively.
Lake Lakar is remarkable. The trees were not
damaged by temperatures as low as −13°C,
and did not suf- fer from snow breakage. A
layer of volcanic ash, rich in iron and
aluminum oxides, was deposited on May 22,
1960, but did not harm the plantation
(Fernandez 1964).
Douglas-fir has been used for rehabilitating
Notofagus antarctica scrubland in the
southern Cordillera of Rio Negro territory
(Gomis 1974) and is considered suitable for
conversion of Berberis scru- bland in the
Patagonian Andes (Seibert 1979). The
planting of 6,200 ha had been projected
according to a national development plan in
the early 1970s (Yakubson 1973). Nothing in
the available literature indicates whether or
not that goal was attained.

Africa
Repeated attempts have been made to
introduce Douglas-fir to Africa, but this has
resulted mostly in failure, as in Kenya (Troup
1932) and Zimbabwe, formerly Rhodesia
(Streets 1962).

Republic of South Africa


Foresters in the Republic of South Africa
initiated trials with both the coastal and
interior varieties of Douglas-fir around the
turn of the 19th century. Troup (1932)
mentioned plantations of Douglas-fir in
silvicultural subdivisions A (mountains of
south- western Cape Province), D (area
between Great Brak and Kromme Rivers), and
H (highlands of eastern Cape Province, Natal,
Transvaal, and eastern and central Orange
Free State). These three subdivisions have
annual precipitation of more than 635 mm.
Legat (1932) considered Douglas-fir as
unsuited for South Africa, and cited as proof
repeated failures of plantations. He asserted
that the intensely dry atmo- sphere of South
Africa, combined with long periods of
drought, is mainly responsible for the failures
of Douglas-fir. Streets (1962) mentioned only
4. Provenance Trials
Richard K. Hermann

P
rovenance trials are designed to study the Washington Cascades on the Snoqualmie NF at
performance of seeds from different stands 610 m; the southern Washington Cascades on the
of trees, which may either be native to Gifford Pinchot (formerly Columbia) NF at 335
their m; the northern Oregon Coast Range on the
place of growth or are introduced from elsewhere Siuslaw NF at 600 m; and the northern Oregon
(Edwards 1956). Usually, provenance trials are Cascades on the Mt. Hood NF at 853 m and 1,402
es- tablished on several sites to assess both the m. In the Mt. Hood NF, six test plantations were
genetic variability of the provenances tested and established at three different sites that differed in
the inter- actions between genotype and altitude. The 1915 and 1916 plantings at the third
environment. Such trials, also referred to as site were destroyed by fire in 1917. In the
common garden studies, have made important Snoqualmie, Gifford Pinchot, and Siuslaw
contributions to the knowledge of genetic National Forests, only one site was used for the
variation in Douglas-fir. Although some were test plantations. Unfortunately, the 1915 plant-
conducted as nursery studies, the majority were ings on the Snoqualmie NF and the 1916
initiated as long-term field trials. The need for plantings on the Siuslaw NF had also to be
long-term trials is clearly expressed by Silen’s abandoned shortly after establishment because of
(1978) statement that, “Meaningful expression of animal damage in the former, and poor survival in
genetic variation can occur early for some traits the latter.
but requires many decades for several that are Notwithstanding some shortcomings in its ex-
commercially important.” perimental design (that is, individual tree
progenies were not replicated and were planted in
Pacific Northwest Studies the same sequence in all plantings), the 1912
Questions arose as to the consequences of seed provenance study has provided some valuable
move- ments with the beginning of artificial answers as to the value of a local seed source in
regeneration in the National Forests of the Pacific regeneration.
Northwest. Because of the economic importance Records of the performance of the 13 seed
of Douglas-fir, a provenance study was initiated sources over a period of 60 years, summarized by
by Raphael Zon in 1912 (Munger and Morris Silen (1965, 1966b, 1978), demonstrate that seed
1936). source and genotype by environment interactions
That first common garden experiment with were mostly of minor consequences at age 17, but
Pseudotsuga menziesii var. menziesii was started had become large at age 60 for both families and
with collections of seed from 120 open-pollinated seed sources. With more than 80% of live trees on
par- ent trees, 15 to 600 years of age, in the fall of all five sites, survival did not differ greatly near
1912 at altitudes that ranged from 90 m to 1,170 the end of the second decade. That pattern had
m. The 13 localities where seed was collected are changed drasti- cally by the end of the sixth
situated between 44°and 49°lat. N in western decade. Then, survival varied from 24% to 64%.
Oregon and Washington (Figure 4.1). Exposure of site turned out to be a significant
Progeny from each parent were planted in 1915 factor in relation to survival. In the Mount Hood
and 1916 as 1−1 stock in four localities: the northern plantation B, at the highest elevation (1,402 m)
and most severe site, progeny from some

73
74 Douglas-fir: The Genus Pseudotsuga
low-elevation races suffered heavy losses during reduced by shorter rotations, by seed movements
the first decade. By age 60, only progeny from involving only minor environmental changes, or
three of the high elevation sources had enough by choice of sheltered sites.”
survivors to form a stand. In plantations at lower Concerning the Wind River test site and the
elevations (335 m and 610 m), but exposed sites, then 80-year-old 1912 provenance study, Silen
mortality was high in maladapted races after age and Olson (1992) concluded that
30. At age 60, several of the 13 sources displayed
• understocking is the primary symptom of
poor stocking and growth on these sites. By maladaptation;
contrast, all but two races had full stocking at that
• inherent growth rates are stable over time; and
age on a sheltered site at 853 m altitude.
Decimation of these two occurred mostly in the • yields are initially related to growth rate but
fifth decade after establishment. become increasingly related to survival.
Perhaps the most important result of the This last point is very important as some
1912 Douglas-fir heredity study was the provenances that had good growth rates for the
demonstration that seed movement can involve first 20 years suf- fered considerable damage and
risks of unaccept- able mortality, but a long time mortality from some extreme climatic event such
may pass before it becomes apparent. As Silen as the November 1956 freeze that damaged and
(1978) stated, however, “The results suggest also killed some mature trees.
that these risks may be
The 1954 Oregon State
49°
University provenance study
Forty-two years after initiation of
Hazel - 275 m Fortson 150 m Darrington - 150 m Snoqualmie - 610 m Granite Falls - 120 m
the 1912 Douglas-fir provenance
study, a second provenance trial
Everett was initiated at Oregon State
Seattle University (Ching and Bever
Tacoma 1960). Like the 1912 study, the
47°
1954 trial was limited to seed
sources from the coastal
Washington
Oc

variety. It included
Race Track - 790 m Wind River - 330 m
Carson - 120 m 16 different provenances from
Vancouver Island to southwestern
Palmer - 610 & 910 m Portland - 90 m Oregon in 16 reciprocal outplant-
Mt Hood - 850 & 1400 m
Siuslaw - 640 m ings of two plantations each, at or
45° Gates - 290 m near the locations of seed
Santiam - 850 & 1170 m Benton - 210 m
collection. In addition, a single
Oregon
plantation that contained the 16
provenances was established in
northern California. Assessments
of the performance of
provenances were made at ages 3
Seed collection site (Ching 1965), 9 (Rowe and Ching
43° Test plantation site
1973), 20 (Ching and Hinz 1978),
and 25 years (White and Ching
04896144 km 1985). The study showed only
small differences at age 25 in the
relative performance of the differ-
125° 123° 121° 119° ent provenances, and none among
Figure 4.1 Seed collection and planting sites of 1912 Douglas-fir heredity study Munger and Morris 1936).
(from
eight low-elevation seed sources.
There was little evidence for inter-
Chapter 4. Provenance Trials 75
growth and highest survival in the archives plantation;
actions between provenances and planting loca- (2) high-
tion. These findings appear to conflict with those
of seedling tests that indicated tight adaptation of
Douglas-fir in the Pacific Northwest to local
environ- ments (Hermann and Lavender 1968,
Campbell and Sorensen 1978, Campbell 1979,
White et al. 1981). A possible explanation for
these seemingly contra- dictory results may be
that a broad genetic mix of families within each
provenance of the 1954 trial tended to reduce, or
even eliminate, provenance x location
interactions. Furthermore, most of the plan-
tations were on fertile sites with mild climates,
and did not experience extreme climatic
disturbances. On such sites, expression of
provenance differences and provenance x location
interactions may require unusual climatic events.
Also in 1954, another provenance trial
commenced
in British Columbia, but it was restricted to just
one planting site at Haney, British Columbia
(Haddock et al. 1967). The trial was established
with commer- cial seed lots representing nine
coastal provenances from British Columbia,
Washington, Oregon, and California, with six
interior provenances from British Columbia,
Montana, and Colorado. Differences in height
growth among the provenances apparent in the
nursery phase remained essentially the same at
age 11. The provenances Ashford, Washington
and Surrey, British Columbia ranked consistently
at the top from ages 2 to 11. Data on ranking at
later ages are not available.
Compared to Haddock et al. (1967), a far more
comprehensive study of geographic variation was
begun in 1957 by Irgens-Møller (1963). He estab-
lished a rangewide source archive containing well
over 600 individual parent trees or stands. The
collec- tions included open-pollinated seed and
occasional live seedlings from 10 western states
and parts of Canada and Mexico.
Growth and survival measurements, coupled
with a complete inventory of the archives planta-
tion in 1989, provided a rare opportunity to assess
geographic patterns of genetic variation in a three-
decades-old plantation of Douglas-fir (Gamble et
al. 1996). Analysis of the plantings made in 1961
and 1987 revealed that (1) local and other low-
elevation coastal provenances had the fastest
altitudinal gradients (10 provenances at 6
elevation provenances from the Cascade sites);
Range also had high survival, but significantly
smaller diameters at breast height; (3) • probe genotype environment interactions by
the response of five standard provenances and
southern interior provenances from Arizona
a local provenance at 22 sites; and
and New Mexico had the smallest diameters
and lowest survival; and (4) the northern
interior provenances from Montana and Idaho
were intermediate in survival and diameter.
These results match a pattern found by Li and
Adams (1989) in a rangewide allozyme study
of Douglas-fir, which led them to distinguish
between three groups, namely coastal,
northern, and southern interior populations.
The striking differences in growth and
survival between coastal and interior
provenances found in the archives plantation
support the distinction of two varieties within
P. menziesii.

The British Columbia Forest Service study


The British Columbia Forest Service initiated
a com- prehensive provenance experiment in
1966 whose principal objective was to develop
biologically sound seed transfer rules for
British Columbia (Schmidt 1973). That year,
the British Columbia Forest Service collected
seed from 76 stands in British Columbia, 7 in
Washington, and 5 in Oregon (Figure 4.2). Of
these 88 seed lots, 26 were included among
the IUFRO collection. Seed lots represented
variety menziesii, except for 10 lots from
interior British Columbia, which belonged to
variety glauca. Each lot comprised a bulked
seed sample from at least 10 parent trees
spaced widely apart to reduce the likelihood
of family relations (Illingworth and St. Pierre
1975).
From 1968 to 1975, provenance plantations
were established with 1−1 bare-root
transplants at 36 lo- cations throughout British
Columbia’s Douglas-fir zone (Illingworth and
St. Pierre 1975). The prov- enance study was
divided into four series (Figure 4.3) in order
to
• screen provenances in five coastal
climatic zones (88 at Sooke and Lookout
Mountain; 77, 58, 52 at Port Renfrew, 77;
58 at Kemano, and 52 at Mud Bay,
respectively;
• test the feasibility of seed transfer across
76 Douglas-fir: The Genus Pseudotsuga

130° 128° 126 ° 124° 122° 120° 118° 116°

British Columbia
54°

Alberta
Kemano
54°

Dean
Raush Kinney Lake
52°
Nusatsum Atnarko
Stuie Lonesome
Noeick
Rivers Tatla Messiter

A Seymour Klinaklini 2000 52°


Otter
B Klinaklini Lac Des Roches
10
Port Hardy Malcolm Southgate
B
50° Jeime E.
Southgate White Brisco
Thurlow A Lake
Nimpkish Kelsey Stella Toba Gramson's D'Arcy
Kaouk Owl
Hernando Quadra Jervis Alta Arrow Lakes
Tahsis Powell Tin Rogers Nahatlatch
Hat
Gold Courtena Tenada Squamish 50°
Forbidden Plateau Denman Alexander
Alberni Tin Pitt Harrison Jaffray
Hat
Nanaimo Haney Hope
Cassidy
Kennedy Chehalis Chilliwack High
Franklin Saltspring Skagit Ymir
48° Cayuse Meade Klinakini Low
Duncan Saturna
Jordan San Juan
Sooke Empress
Darrington
Sequim Granite
Hoh Falls 48°

Shelton Washington

46°
Alder Lake

Castle Rock

46°

Grande Ronde

44°

Waldport

44°

Coquille
Olalla
0 30 60 90 mi Oregon
Brookings
0 48 100 144
km

42°
128° 126 ° 124° 122° 120° 118° 116°
Figure 4.2 Collection sites for 1966 British Columbia Forest Service provenance study (from Schmidt 1973).
Chapter 4. Provenance Trials 77

130°
128° 126 ° 124° 122°

54°

British Columbia

Kemano

52°

Bella Coola

Klinaklini
50°
Coal Harbour Toba
Kelsey Bay Jervis Pemberton Low Lois LakePowell R. High
Jeune
Roberts Spuzzum
Powell R. Low
Series 0.01 Landing Lake Quadra Island
Oyster River Piggot Creek Harrison
0.02 Gold River Texada Island Dove CreekChehalis
48° 0.03 Lookout Mtn. Coquhalla
0.04
Kennedy Lake Fleet River Port RenfrewMt. Prevost
Saltspring Valentine Mtn. Sooke High
03060 90 mi Sooke (Muar Cr.)
Figure 4.3 Sites of a series of four provenance plantations established 1968 to 1975 (from Illingworth and St. Pierre 1975).
048100 144 km

• compare three representative coastal and four within that region, the performance of local
representative inland provenances at three
popula- tions was inferior to that of introduced
high-elevation sites on Vancouver Island.
populations. The sites where local provenances
In his summary of 6-year results, Illingworth grew best were located in the Coastal Western
(1978) stated that they confirmed the remarkable Hemlock subzone a (CWHa),1 the drier of the two
variability of Douglas-fir populations, not only in CWH subzones. The most vigorous populations
their ability to survive and grow on a particular originated from within the Coastal Western
site, but also in their responses to different Hemlock zone (CWH), and they invariably
environments. The data did not demonstrate a outgrew populations from the Coastal Douglas-
pattern of variation in relation to geographic fir zone (CDF), especially when planted in that
variables. zone. Geographically, the CDF encompasses the
Trees from the coast-interior transition zone islands and lowlands rimming the Strait of
and north-mainland populations were generally Georgia
shorter, suggesting a relationship between latitude and Puget Sound.
or eleva- tion of population origin. However, in The results also indicated that populations can
the extensive region between the Coast and be transferred up or down in elevation without
Cascade Ranges and the mountains of Vancouver seri- ous consequences within the CWH zone. But
Island and the Olympic in the high-elevation Mountain Hemlock zone
(MH), or
Peninsula, there was no clear relationship between
vigor and these variables. Moreover, at most sites 1. Zonal classification by Krajina (1969).
78 Douglas-fir: The Genus Pseudotsuga
in the mainland valleys of the central coast, high- provenance means in-
elevation or north-coast populations appeared har-
diest. Although they were less vigorous, they are
preferable for use in these locations.
At plantation age 20, the patterns observed had
not changed from those observed at earlier ages
(Ying 1990). The provenances belonging to variety
menziesii continued to exhibit broad regional
differentiation corresponding to the major
climatic regions of their origin. Provenances from
the moist maritime zone maintained their
superiority over those from the dry lowland and
the cold coast-interior transition zone. Elevational
differentiation was not apparent, however, and
regional differences accounted for about 50% of
the among-provenance variation.
The results at age 20 appear to lend support to
Illingworth (1978), who inferred tentatively from
the 6-year results that, “Douglas-fir varies
ecotypi- cally throughout a considerable part of its
coastal distribution, a primary ecotype being a
zone of op- timum growth.” The substantial
variation between provenances in that zone still
awaits an explanation, however.
The University of British Columbia trial
Oscar Sziklai, Professor of Forest Genetics at the
University of British Columbia, established a
Douglas- fir provenance test in 1971 at the UBC
Research Forest at Haney, lat 49°16’ N, long
122°34’ W, elevation 145 m, with 102
provenances from the 1966 and 1968 IUFRO seed
collections. Seedlings were planted as 1+1 plugs
in 1971. The objectives of the study were to
estimate the degree of genetic variation among
and within provenances under a specific local
condi- tion, the magnitude of genetic control over
growth characteristics, and the strength of
juvenile-mature correlations (Fashler et al. 1987).
Height measurements of trees in 348 families
from 11 British Columbia, 16 Washington, 15
Oregon, and
6 California provenances were made in 1972,
1975, and 1978, and then analyzed by grouping
prov- enances according to four seed zones. Three
of the seed zones represented coastal and the
fourth zone interior provenances. Results of the
height measure- ments demonstrated significant
variation in height growth among and within seed
zones. The observed large range in seed zone and
dicated considerable genetic variation. This 42°00’ and 48°15’ N. The differen- tial
finding suggested that substantial gains were response to soil and air temperatures by these
achievable by selecting the most desirable populations was attributed to differences in
provenances. Ranking of provenances their genetic makeup.
according to mean 1975 and 1978 total height
for all provenances showed little change in the
ranking of the best 25% of the provenances,
which suggested good reliability in the
selection of the best provenances at age 5.
Sziklai (1990) wrote that, “Even after 16
growing seasons only the extreme families
show a certain consistent performance but the
large majority of the families standing
between the best and poorest performers still
alter their position year after year.” Of the five
tallest provenances after 16 growing sea- sons
at Haney, four originated from Washington
and only one from British Columbia. The
most southerly of the four Washington
provenances came from Naselle (2°54’ south
of Haney). As to the question of which
provenances are best suited to a specific site,
and how long a wait is necessary to be certain
of the right choice, Sziklai (1990) concluded
that, after 16 growing seasons, the question
can be answered with more certainty than
before.
Short-term seedling tests
Short-term seedling tests of coastal Douglas-
fir in outdoor nurseries, greenhouses, and
growth cham- bers typically involve a detailed
assessment of adap- tive traits in the first one
to three years, whereas long-term field tests
measure overall tree growth over a much
longer period of time (Howe et al. 2006.)
Seedling tests in common garden
environments and allozyme studies have
indicated genetic varia- tion of the variety
menziesii on a macrogeographic scale that
appears to follow gradients (clines) de-
scribed by latitude, elevation, and distance
from the ocean. Some tests have shown that
microgeographic factors—aspect for instance
—may modify patterns
of variation in quantitative traits.
A steep east-west genetic gradient of
coastal Douglas-fir in its central range was
reported by Campbell and Sorensen (1978).
They sampled 40 populations from elevations
ranging from 6 to 1,400 m between lat
Chapter 4. Provenance Trials 79
450 to 1,500 m
An investigation of microgeographic variation
of Douglas-fir populations was conducted by
Campbell (1979). He estimated genetic values of
193 parent trees throughout a 6,100 ha watershed
in west-central Oregon from progeny grown in a
common gar- den through three growing seasons.
Variation was partitioned into components
attributable to parent tree location and differences
among trees within locations. Campbell described
the large within- location variation as being
homogenous within the watershed because
variation of traits among trees within locations did
not differ among locations. By contrast, virtually
all variation could be accounted for by the
location of parent trees. The variation pat- terns in
the watershed suggest a three-dimensional cline in
which trait values are a function of elevation and
north-south or east-west location, the function
varying somewhat for each trait. He concluded
that topoclinal variation in traits, as well as the
large within-location variation, are the
consequences of high selection intensities in the
seedling stage, the former to selection by average
environmental dif- ferences along gradients, the
latter to microenviron- mental heterogeneity.
That the majority of seedling tests in common
garden environments concerned with the coastal
variety has focused on Douglas-fir from southwest
Oregon and northwest California seems to reflect
the fact that these are regions of great
vegetational, edaphic, topographic, and climatic
complexity. In some areas in these two regions,
Douglas-fir dis- plays many of the characteristics
of a species near its adaptional limits—that is,
difficult regeneration, a distribution influenced by
topography, and all- aged stands of a species that,
in the central parts of its range, is commonly
represented by even-aged stands. To maintain
fitness in such a heterogenous environment,
Douglas-fir, in the opinion of Campbell (1987), has
generated many genotypes and a large amount of
genetic variability.
A common garden study by Hermann and
Lavender (1968) in southwest Oregon provides
ev- idence that microgeographic factors may
modify patterns of phenotypic variation along
individual mountain slopes. They collected open-
pollinated cones from south- and north-facing
aspects at 150- m intervals along a transect from
42°20’ and 42°40’ N provided evidence of steeper
elevation on the western slope of the Cascade genetic clines (eleva- tional and latitudinal) along
Range between lat 42°00’ and 43°12’ N. The the coast than farther inland. The investigation of
progeny were grown for two years in two seed and seedling traits was based on a collection
nurseries, one in south- ern, and the other in of open-pollinated seed
northern Oregon. They were also grown in
growth rooms under six combinations of
thermoperiod and photoperiod. Trends of
varia- tion in traits with elevation were
shown to depend partly on slope and aspect.
A geographically broader study (White et
al. 1981), intended to represent the range of
sites on which Douglas-fir grows in southwest
Oregon, in- volved samples of wind-
pollinated seed from two parent trees at 36
locations between lat 42°00’ and 43°12’ N.
Elevation of collection sites ranged from 475
to 1,630 m, 61 to 162 km inland from the
Pacific Ocean. Seedlings were grown in three
test environ- ments (growth room, nursery,
and greenhouse) to assess environmental
influences on genetic variation of first-year
height growth. In all three test environ- ments,
mean first-year population height growth
correlated most strongly with elevation of
place of origin. Seed collected at higher
elevations—that is farther inland—produced
shorter seedlings.
Subsequently, Campbell (1986) described pat-
terns of genetic variation based on 135 parent
trees from 80 locations in southwest Oregon.
The area he sampled is nearly square with
west and east boundaries about 60 and 190 km
from the Pacific Ocean. The southern
boundary is on the border with California, and
the northern boundary is along the 43°N
parallel. Campbell estimated genetic values
for 13 traits from the open-pollinated progeny
grown in nursery beds. The pronounced east-
west gradients shown in this experiment
followed trends of steep east-west gradients
farther north at high elevations in western
Oregon and Washington demonstrated by his
earlier work (Campbell and Sorensen 1978),
and those reported by Griffin (1978) for the
Coastal Ranges of northwest California, and
by Sorensen (1983) for the western Siskiyou
Mountains.
Sorensen’s (1983) study of genetic differentiation
of Douglas-fir in the lower Rogue River
watershed of southwest Oregon between lat
80 Douglas-fir: The Genus Pseudotsuga
made at four elevations between 150 and 1,065 m ferentiated sub-populations expressing a comparable
on west- and east-facing slopes on the first two range of character combinations.”
ridges inland from the Pacific Ocean. Trait
differences were generally greater between the
west and east aspects of the coastal ridge than
between the two aspects of the inland ridge.
This pattern of variation appeared to be
determined by adaptation to local moisture and
temperature regimes.
A comparison of the genetic structure of
Douglas- fir from different habitats in southwest
Oregon in a common garden seedling study
(Hamlin 1990) included eight populations, two
each from four major conifer zones: the Tsuga
heterophylla zones in the Coast Range and the
western Cascade Range, the mixed-evergreen
zone in the Siskiyou Mountains, and the mixed-
conifer zone in the South Cascades- Klamath
Range. Measurements of 19 traits over two
growing seasons indicated that differences in
genetic structure did not vary randomly, but were
associated with the extent of habitat divergence.
A nursery study of 18 wind-pollinated families
from the North Coast Ranges and Klamath
Mountains of northwestern California (Griffin and
Ching 1977, Griffin 1978) included progeny from
85 stands at 9 locations within an area extending
from lat 37°08’ N, long 122°11’ W to lat 41°47’,
long 124°00’ W. At each location, the altitudinal
range of Douglas-fir had been sampled by spacing
seed collections at 76-m intervals. All assessed
traits varied geographically. With the exception of
time of budburst, the most significant contrast was
between populations from the coastal fog belt and
those from the interior ranges. Coastal seed was
smaller and germinated more slowly than seed
from the interior ranges. Coastal seedlings had
fewer cotelydons but greater epicotyl growth,
grew for a longer period before setting buds,
showed less capacity to set buds in response to
moisture stress, and were less cold hardy than
those from the interior ranges. But in spite of
broad similarities in variation patterns,
distribution of variation among sampling levels
was not the same for all traits. Griffin and Ching
concluded: “The most satisfactory concept of the
northern California Douglas-fir population is that
of a single gene pool within which a complex
spectrum of selection pressure gradients have dif-
A later study by Kitzmiller (1990) of breeding zones within 13 geographically designated areas called
breeding units. Breeding zones are elevational bands, each of which
genetic varia- tion among northwestern spans an altitudinal range of < 300 m and is generally smaller than
California Douglas-fir was based on height 60,000 ha. They were established in regional Douglas-fir tree
improvement programs (Silen and Wheat 1979).
growth of 675 open-pollinated families
planted in 1980 in native locales. The in-
vestigation included also allozyme analyses of
seed from 315 parent trees. His findings
indicated the association of a substantial
amount of genetic varia- tion with latitude,
longitude, and elevation of seed source. Seed
weight patterns followed a northwest to
southeast cline of increasing seed weight. The
main factors associated with changes of seed
weight were distance from the ocean and
mean annual precipita- tion. The lightest seeds
came from an area 13–16 km distant from the
ocean and the heaviest seeds from 40–48 km
inland.
Specific gravity of wood exhibited a different
pattern. Apparently wood densities of coastal
and Klamath Mountains populations are
similar at similar elevations, but major
differences occur lo- cally between low and
high elevation populations. Allozyme
patterns, just like those of seed weight,
change from northwest to southeast, and
with el- evation. They separated trees into 8
groups, which formed four geographical
zones in a longitudinal direction: coastal,
west central, east central, and east- ern,
subdivided further into two elevational bands.
In marked contrast to variation in the
quanti- tative characters found in seedling
common gar- den studies, results of allozyme
studies of coastal Douglas-fir, except one by
Kitzmiller (1990), did not indicate
associations between allozyme variation and
environmental variables. Merkle and Adams
(1987) studied the distribution of allozyme
diversity among 22 breeding zones in
southwestern Oregon based on the
electrophoretic analysis of haploid
megagametophytes from 1,230 parent trees.
These 22 breeding zones provide elevational
transects in each of seven breeding units,2
west-east from the coast inland, and north-
south along the coast and inland. Climate
differs widely along these transects. Mean
annual precipitation decreases rapidly west to

2. Southwest Oregon has been subdivided into 38 Douglas-fir


Chapter 4. Provenance Trials 81
the pattern of varia- tion for seven different
east, and minimum winter temperatures quantitative traits. Only 7% of the total genetic
decrease with distance east from the coast and diversity was attributed to dif-
with increas- ing elevation. Merkle and Adams
reported that less than 1% of the variation could
be attributed to differ- ences among zones. The
lack of a relation between allozyme variation and
geographic variables was in striking contrast to a
companion common garden study that, although
based on the same material, re- vealed
significant clinal patterns of genetic variation for
quantitative traits (Loopstra 1984, Loopstra and
Adams 1989). Use of a different technique,
multivari- ate analysis of allozyme variation
patterns to describe the distribution of genotypic
variation among and within these 22 southwest
Oregon breeding zones also failed to provide
evidence that allozyme varia- tion is adaptive in
the coastal Douglas-fir breeding zones studied
(Merkle et al. 1988).
In another attempt to ascertain whether a
similar
adaptive differentiation for allozyme variation
exists in the same heterogenous environments as
has been shown for quantitative traits, stands of
Douglas- fir on nearby south- and north-facing
slopes were genetically compared in two of the
breeding zones in southwest Oregon (Moran and
Adams 1989). Samples for the study consisted of
60 trees from each of 12 stands. Dormant buds (as
well as needle tissue in one stand) were used for
enzyme analyses. These analyses indicated that
the proportion of to- tal genetic diversity that
resulted from differences among stands was only
1.8%, a percentage close to that attributed to
differences in the distribution of allozyme
diversity among 22 breeding zones in southwest
Oregon (Merkle and Adams 1987). The obvious
conclusion is that abrupt changes in slope aspect
or steep elevational gradients within breed- ing
zones in southwest Oregon appear to have little
influence on the genetic structure of Douglas-fir
as shown by allozymes.
Results of the allozyme studies in southwest
Oregon agree with those of an earlier study in
south- west British Columbia (El-Kassaby and
Sziklai 1982). That investigation described genic
patterns at 27 different allozyme loci in a natural
stand of Douglas- fir along an elevational transect
divided into four elevational segments, as well as
Colorado grew two or three times as fast as those
ferentiation among the elevational segments; from popula- tions belonging to the northern
the remaining 93% resided within segments. subgroup of interior Douglas-fir. But the former
The traits studied showed the same general incurred more frost damage and winter injury
trend of variation in the different elevational than the latter. By age
segments ranging from 94% to 100%.

Common garden studies: variety glauca


Except for the rangewide source archive
plantation established in 1957 by Irgens-
Møller (Gamble et al. 1996), reports of
observations extending over sev- eral decades
have not been published for any of the
common garden studies of variety glauca in
North America. Wright et al. (1971) initiated
a comprehen- sive study in 1961 to determine
geographic variation patterns in interior
populations of Douglas-fir, and to identify
suitable provenances for Christmas tree and
ornamental use in the north-central states. The
study included 74 provenances belonging to
the northern subgroup and 33 provenances
from the southern subgroup of inland
Douglas-fir, as well as 21 provenances of the
variety menziesii. The tall- est seedlings were
outplanted as 1-2 stock in 1965, the remaining
ones as 1-2-2 stock in 1967, in three locations
in southern Michigan. Five-year results
indicated that Arizona and New Mexico
provenances grew rapidly enough to be
harvested as Christmas trees 6 to 8 years after
planting, but those from the central and
northern Rocky mountains grew con-
siderably slower. By contrast, the pattern for
cold and frost hardiness was just the opposite:
the slower growing northern provenances
were hardier than the faster growing southern
provenances.
A provenance test begun 1965 in Plattsmouth
in eastern Nebraska (Read and Sprackling
1976) included seedlings of 26 provenances
from the north- ern subgroup and 15
provenances from the southern subgroup of
inland Douglas-fir. Seedlings came from the
collection made for the Michigan study by
Wright et al. (1971). Data that covered
performance for the first 11 years in the
plantation showed that growth rates of
provenances were inversely related to their
latitude of origin. Progeny of populations
from Arizona, New Mexico, and southern
82 Douglas-fir: The Genus Pseudotsuga
22, this pattern had essentially remained the same belonged to the northern subgroup.
(Van Haverbeke 1987). Maladaptation of the southern provenances to
In 1966, the British Columbia Forest Service the Trinity Valley environment may account for
col- lected seed lots from 64 stands distributed the slow growth at that site. Jaquish (1990)
over the entire natural range of variety glauca concluded that, “On mild, low elevation sites
for a prov- enance trial. Each seed lot consisted of within the ICH biogeoclimatic zone of the
bulked, open- pollinated seed from a minimum of southern Interior, seedlings from most
10 dominant or codominant trees per population. provenances belonging to the northern race can
The trial had as its objective to determine broad be established and expected to survive; however,
patterns of geographic variation in interior major differences in productivity can be
Douglas-fir and to generate information about expected among provenances. To maximize
growth and adaptability for a broad spectrum of productivity in planta- tions on these sites,
var. glauca populations in the southern interior provenances originating from the Shuswap
of British Columbia (Jaquish 1990). The seeds Lake/North Thompson River area are
were sown in spring 1972 at the Cowichan Lake recommended for planting.”
nursery. The resulting seedlings were planted in A series of seedling common garden studies by
fall 1975 in a replicated experi- ment at the Rehfeldt (1974a, 1974b, 1978, 1979a, 1979b, 1982,
Trinity Valley Tree Breeding site near Enderby in 1983a, 1988) have provided by far the best insight
southern interior British Columbia. The site is at into the reasons for the variability of quantitative
600 m elevation in the cool, moist Interior Cedar traits in the northern subgroup of variety glauca.
Hemlock (ICH) biogeoclimatic zone.3 After 10 These experiments, which compared populations
growing seasons, survival for the entire planta- throughout the range of the northern subgroup of
tion was 87%. Survival of provenances ranged va- riety glauca, showed steep adaptive clines for
from 0 and 1% for the two Mexican provenances inland Douglas-fir in northern Idaho and western
to 100% for five provenances that belonged to the Montana. Rehfeldt (1989) summarized patterns of
northern genetic variation in interior Douglas-fir across
subgroup of var. glauca. 250,000 km2 of forested lands in the northern
Provenances varied tremendously in height at Rocky Mountains, based on results from four of
age 13 from seed, with a range from 130 to 472 his common garden studies. He used data from
cm. Most provenances—that is 21 out of 23—that 228 populations included in these tests to show
belonged to the southern subgroup ranked in the the extent of genetic variation within four
lower half, below 300 cm, of the height range. Of physiographic provinces; in northern
the 13 tallest provenances, 7 originated from the Washington and northern Idaho (Rehfeldt 1979b);
low to middle elevations of the interior wet belt northwestern Montana (Rehfeldt 1982); central
(ICH) zone, and 6 came from the dry subzones of Idaho (Rehfeldt 1983a); and southwestern
the east slopes of the Coast mountains and Montana and Idaho, near the Continental Divide
Cascade range (Coast Interior Transition zone). (Rehfeldt 1988). For each of these studies, seeds
Provenances from the Shuswap Lake and North were collect- ed from open-pollinated trees in
Thompson River areas were the tallest among the natural stands. Their progeny were grown for
provenances from the ICH zone. These findings three years in two of the same common gardens
indicated strong racial differences in height at elevations of 762 m and 1, 524 m near Priest
growth of interior Douglas-fir and, in particular, River in northern Idaho. Comparisons of the
between its northern and southern subgroup in seedling populations were based on several traits
height growth patterns. But the results are at that reflected growth, development, and frost
variance with those from provenance tests by tolerance. Rehfeldt used third-year height to index
Wright et al. (1971) and Van Haverbeke (1987), adaptive differentiation because this vari- able
who found that provenances from the southern was the only trait common to all studies but was
also the most strongly correlated with other traits.
subgroup grew much more rapidly than those that
Populations from elevationally or geographically
3. Zonal classification by Pojar et al. (1987). mild sites were tall but had low freezing tolerance,
and populations from harsh sites were short
and
Chapter 4. Provenance Trials 83
opportunity for institutions
cold hardy. Rehfeldt (1991) constructed a model
of genetic variation from the data provided by
these studies that produced elevational clines of
the same shape for all locations. The slope of the
elevational clines varied geographically, however.
For instance, the elevational cline is steepest in
northern Idaho and northeastern Washington,
where Douglas-fir grows at the lowest elevations
but is nearly flat in southwestern Montana where
the species occurs only at high elevations.
Models are invariably subject to errors and re-
quire verification with independent data. A strong
validation of this model of genetic variation
resulted from a study that attempted to answer the
ques- tion of how much variation in site index is
associ- ated with genetic variation. Monserud and
Rehfeldt (1990) correlated the genetic variability
predicted by Rehfeldt’s (1989) model with the
mean 50-year height of three trees in each of 135
natural stands in northern Idaho and western
Montana. The genetic variability predicted by the
model accounted for 42% of the variation in 50-
year dominant height among these 135 stands.
Rehfeldt (1991) concluded that the variability of
the northern subgroup of interior Douglas-fir has
been brought about by environmen- tal selection
to produce populations physiologically adapted to
specific segments of the various environ- mental
gradients. Nonetheless, substantial genetic
variation exists within populations.

The IUFRO International


Douglas-fir Provenance Study
The introduction of P. menziesii into the
temperate zones of the southern and northern
hemisphere, out- side of North America, led to
numerous provenance trials of the species in
various countries during the first half of the 20th
century. Most of these trials suffered from two
shortcomings, however. One was lack of precise
information about the geographic location of the
seed sources and the method of seed collection,
and the other was very limited coverage of the
natural range of Douglas-fir.
The project of a range-wide seed collection
initi- ated by the former Section 22 of the
International Union of Forest Research
Organizations (IUFRO) in 1965 provided an
Baron von Fürstenberg established six test plan-
not only to obtain seed of exactly known tations between 1904 and 1911 on his estates in
origin but also the prospect of comparing
results later with others who conducted
provenance tests with the same material. The
IUFRO seed collection was or- ganized by H.
Barner of the Danish State Forestry Tree
Improvement Station in Humlebaek,
Denmark. A total of 326 kg of Douglas-fir
seed was collected from 182 stands in the
years 1966/68/69/70. To fa- cilitate revisits
for additional collections, the exact location of
each of the 182 seed sources was marked on a
large-scale map (Fletcher and Barner 1987).
To eliminate any inbreeding effects, cones
were har- vested in each stand from 15 to 20
dominant and codominant trees distributed
equally throughout the stand within a distance
of about 100 m from each other. Seeds were
kept separate for each tree, so that single-tree
samples could be provided for institutions that
requested them. The remaining seeds from
each provenance were bulked before dis-
tribution to participants in the IUFRO
International Douglas-fir provenance study.
By 1973, samples had been distributed to 45
institutions in 30 countries (Barner 1973) in
both the northern and southern hemispheres.
Published findings that pertain to the IUFRO
International Douglas-fir provenance study
are covered after the discussion of earlier
European provenance trials.
Pre-IUFRO European Studies
Douglas-fir was introduced into Europe in the
19th century with seed of the variety
menziesii, but the location of the mother
stands is unknown. Trees from these early
plantings adapted well to the en- vironment
and often surpassed the growth of native
European conifers. The early plantings had
been on a small scale and little thought had
been given to the origin of seed. That changed
as some European foresters advocated larger
use of Douglas-fir in forest plantings and
realized the need for more knowledge about
suitable seed sources. Consequently, numer-
ous provenance tests were initiated in Europe
in the 20th century.
German provenance trials
The 1904–1911 Fürstenberg trial
84 Douglas-fir: The Genus Pseudotsuga
Westphalia. Kanzow (1937) referred to that two from the Snoqualmie area in Washington. That
undertak- ing as the “caesia provenance trial of only two western Washington provenances
Fürstenberg,” although it would hardly qualify as
a provenance test by present-day standards. The
Fürstenberg trial is included here, however,
because it represents the first recorded attempt to
test the growth of Douglas- fir from the transition
zone between the coastal and inland variety in
Germany. Apparently, some of the trees in his test
plantations originated from seed Fürstenberg had
collected during his travels in British Columbia in
1903 and 1904. He did not indicate as to how
many stands his collections rep- resented, merely
that they came from areas in the vicinity of Field
and Quesnel, British Columbia (Fürstenberg
1923).
Fürstenberg did not plant any coastal
provenanc-
es for comparison. But Kanzow compared mean
height of trees in Fürstenberg’s plantations with
his yield tables (Kanzow 1937) for coastal
Douglas-fir. He found that growth of
Fürstenberg’s trees was intermediate between his
site class I and II for coastal Douglas-fir, at least
until age 33. Information about the subsequent
development of the Fürstenberg plantations was
not found.
The 1910 Schwappach trial
Shipments of seed of inland Douglas-fir to
Germany by the U.S. Bureau of Forestry from
1891 to1895 led eventually to the initiation of a
provenance trial by the former Prussian Forest
Experiment Station. Adam Schwappach, its
director from 1899 to 1925, had noticed what he
conceived to be geographic variability in
progeny from these seed imports and decided to
investigate the importance of seed source for
future procurements of seed (Schwappach 1907).
He established a provenance test at Chorin,
Brandenburg, in spring 1910 with seed from the
1909 cone crop obtained through the help of
Raphael Zon of the US Bureau of Forestry
(Schwappach 1914). The first good seed year after
the crop failures in 1907 and 1908 was in 1909
(Schwappach 1909). The seed collection consisted
of 19 provenances, 12 from the northern and 2
from the southern subgroup of variety glauca,
and 5 of the variety menziesii. Three of the
coastal provenances were from California and
were included in the collection sent to Assessments made during that span of time were
Schwappach is surprising in view of the fact published by Münch (1923, 1924, 1928), Kanzow
that Zon had told Schwappach (1911) that (1937), Boiselle (1953), Rohmeder
seed from the western slopes of the Cascades (1956), Stimm (1995), and Stimm and Dong
between lat 45°and 50°N would be best suited (2001). The earlier assessments, as well as the last
for use in Germany. at age 90 from seed, showed that Snoqualmie, the
Trees in the Chorin plantation were measured only var.
at
ages 5, 18, 25, and 44 from seed (Schwappach
1914;
Kanzow 1936, 1937; Flöhr 1954). Growth
ranking remained essentially the same from
age 5 to 44. The two Snoqualmie provenances
were the best performers and had reached
heights of 27 and 26 m and diameters breast
height of 31 and 29 cm, respec- tively. Height
of the three California provenances was 4 to 6
m less, and that of the surviving inland
provenances was from 8 to 14 m lower.
Infection with Rhabdocline pseudotsugae,
first ob- served in 1930, caused serious
damage to the inland provenances in the
following year but remained without
noticeable effects on the coastal provenances
(Liese 1932, 1935, 1936). Flöhr (1954)
reported that, by 1953, 5 of the 14 inland
provenances had been practically eliminated
and the remaining 9 showed low vigor.
The 1912 Kaiserslautern trial
Ernst Münch, Professor of Forest Botany at
Ludwig- Maximilian University in Munich,
established a par- allel test with 10 of the 19
provenances in the Chorin trial, during his
tenure as supervisor of the former
Kaiserslautern-Ost Forest District in
Rhineland- Palatinate. Schwappach had sent
Münch 8,000 two- year-old seedlings which
were planted in spring of 1912 in the
northern Pfälzerwald mountain range. The
test plantation is located at lat. 49°25’N,
long. 7°40’E. Seedlings belonged to one var.
menziesii prov- enance from northwest
Washington and to nine var. glauca
provenances, four from the variety’s southern
subgroup and five from its northern
subgroup.
Performance of provenances in this trial
has been followed for 88 years, which is
probably the lon- gest period of observation
for any of the European provenance tests.
Chapter 4. Provenance Trials 85
provenances, 15 from the Pacific Northwest, 1 from
menziesii provenance in the trial, was consistently Colorado,
the best performer throughout the length of the
trial. The total cumulative yield at age 90 was 1,
958 m3/ ha with bark. The MAI at that age was
21.8 m3/ha. The volume production of the three
surviving var. glauca provenances reached not
even half of that of the var. menziesii provenance.
The provenance Bitterroot from western Montana,
which ranked second in volume production, had
only a total cu- mulative yield of 847 m3/ha and a
MAI of 9.4 m3/ha. Trees remained free of diseases
until 1932 when Münch observed infection with
Rhabdocline pseu- dotsugae on two of the
provenances. In subsequent years, the disease
spread to all inland provenances. Phaeocryptopus
gaeumannii appeared in the plan- tation in 1939.
By 1967, nearly all trees of six of the var. glauca
provenances had succumbed to the needle cast
fungi. Infection with both Adelopus and
Phaeocryptopus gaeumannii was first noticed in
trees of the provenance Snoqualmie in 1951. The
infec- tions were never so severe as to have a
major effect
on growth.
The 1930s trials
In spite of the shortcomings of the first German
provenance tests of Douglas-fir (that is, very
limited coverage of the species’ range and lack of
replica- tions), they yielded one important result.
The tests demonstrated that var. glauca is poorly
suited for planting under most forest conditions in
Germany because of its slow growth and its
susceptibility to severe infection with
Rhabdocline pseudotsugae. The realization that
var. menziesii was inadequately represented in the
initial provenance tests led to ini- tiating several
provenance tests mainly with coastal Douglas-fir
in the 1930s. The good Douglas-fir cone crop of
1929 in many parts of the Pacific Northwest,
which made seed readily available, seems to
account for the fact that these tests were begun at
about the same time.

The 1932–1933 Wiedemann trials


The most comprehensive of these trials was due
to the initiative of Eilhard Wiedemann, Director
of the former Prussian Forest Experiment
Station (Schober 1954). His test included 19
** Destroyed by fire
and 3 from German stands of coastal Douglas- *** Fate unknown
fir of unknown origin. Some of the American
provenances had identical geographic
designations, but they may have differed
somewhat by elevation of seed source. The
seed was bought in 1932 from the Long- Bell
Lumber Company in Longview, Washington.
Wiedemann had secured the cooperation of 12
for- est districts in various parts of Germany
and of the Danish and Hungarian Forest
Experiment Stations to establish test
plantations to compare the per- formance of
provenances under different growing
conditions. He divided the 19 provenances
into two series with the assignment of 8 of the
provenances to 7 German forest districts, and
11 of the provenances to 5 German forest
districts and to the Danish and Hungarian
cooperators (Tables 4.1 and 4.2).
Wiedemann’s choice of locations for the test plan-
tations represented a cross section of the
major cli- matic provinces of Germany.
Although 10 of the test plantations have been
destroyed or have met with an unknown fate,
those that have remained avail- able for
observation are in regions of contrasting
climates. The Danish and Rosengarten
plantations are on low-elevation sites with a
pronounced oce- anic climate. The Braunlage
plantation represents

Table 4.1 Provenances and test locations in the 1932 to 1933


Wiedemann trials (modified from Schober 1954), Series I.
Provenance Elevation (m)
Salmon Arm, British Columbia 900
Stella, Washington 60
Ryderwood, Washington 250
Spirit Lake, Washington 1100
St. Helens, Washington 2000
Lebanon, Oregon 550
Sweet Home, Oregon 900
Kleinengstigen, Würtemberg ?
Plantation location
Korpele, Poland (formerly East Prussia)* 160
Eberswalde, Brandenburg** 30
Freienwalde (section 151), Brandenburg* 95
Rosengarten, Lower Saxony (formerly Harburg) 40
Namslau, Lower Silesia, Poland*** 121
Schönlanke, Pommerania*** 75
Kirchzarten, Black Forest 1300
Braunlage 600
* Destroyed by frost
86 Douglas-fir: The Genus Pseudotsuga
Table 4.2 Provenances and test locations in the 1932 to 1933 harsh high-
Wiedemann trials (modified from Schober 1954), Series II.

Provenance location Elevation


Kamloops, British Columbia 800
Salmon Arm, British Columbia 100
Elma, Washington 60
Ryderwood, Washington 200
Spirit Lake, Washington 1100
Snoqualmie, Washington 1650
Lebanon, Oregon 500
Cascadia, Oregon 900
Gaildorf, Württemberg 329
Lauterbach, Hesse 370
Plantation location
Freienwalde Brandenburg (section 171) 95
Krasiejów, Upper Silesia*** 189
Pfeil, East Prussia* 132
Braunlage, Lower Saxony (formerly 560
Hohegeiss)
Giessen, Hesse* 200
Nødebo, Denmark 9
Kompedal, Denmark 65
Hungary***
* Destroyed by frost
** Destroyed by fire
*** Fate unknown

a medium elevation site in the transition zone


from an oceanic to a continental climate. The
Freienwalde plantation is subject to the
continental climate of the northeast German
plains, and the one at Kirchzarten has the harsh
climate typical of high elevations in the mountains
of southwestern Germany.
Kanzow (1937) reported on the initial phases
of the plantations in the Wiedemann study.
Schober (1954) reviewed available information on
plan- tations of Series I and II at age 24 from seed
and analyzed in detail development of the
plantations at Kirchzarten and Braunlage. The
following year Schober and Meyer (1955)
reported on the plantation at Rosengarten; it had
been presumed a war loss but it and the papers
containing the layout of plots were relocated in
the early 1950s. Erteld (1948), Dittmar (1954),
and Dittmar and Knapp (1967) provided ac-
counts of the state of the Freienwalde
compartment 171 plantation.
The best growth performers at age 24 from
seed were provenances from low and medium
elevations of the western slope of the Washington
Cascades, except on the sites with a continental or
elevation climate. Provenances from consequence was heavy dam- age by early and
sources in the Cascades above 1,000 m had late frosts. At the Rosengarten site, seedlings were
excellent growth at the 1,300 m site in the planted under the shelter of 85- year old Scots
Black Forest (Kirchzarten District), but they pines that prevented frost damage. Subsequently,
performed marginally elsewhere, as did the however, competition from the pine
Oregon provenances in all test plantations.
The British Columbia provenances from the
northern subgroup of var. glauca grew
rather poorly in the maritime climate of the
coastal region of northwest Germany but
they exhibited good growth in the
continental climate of the northeast German
plains. That rate of growth is shown by
measurements at age 37 in the Freienwalde
plantation where the two inland provenances
from British Columbia retained their leading
position in total volume growth. The
Colorado provenance, however, was a
complete failure at Freienwalde, as in all
other test planta- tions. Performance of the
progeny from German Douglas-fir stands
was equal to that of the best American
provenances.
The Kirchzarten plantation, elevation 1,300 m
was remeasured in 1981 at age 52. The low-
elevation provenances from the Washington
Coast Range showed the best growth, but the
superiority of the high-elevation provenances
from the Washington Cascades observed in
1952 had not been maintained. Estimates
based on the 1956 yield tables for Douglas- fir
by Schober indicated an M.A.I. at age 75 of 2
to 5 m3/ha. That estimate pointed to poor
prospects for Douglas-fir at high elevation
sites in the Black Forest (Kenk and Thren
1984b).
The Wiedemann study suffered several
short- comings quite common in early
provenance studies. Provenances were not
replicated in the test planta- tions. The
condition of seedlings at outplanting in the
field appears to have differed significantly
between plantations. First- and second-year
survival was highest, 88%, at Rosengarten,
the site nearest to the Halstenbeck nursery
where all the seedlings for the experiment had
been raised. It was lowest, 55%, at the sites in
East-Prussia and Upper Silesia farthest away
from the Halstenbeck nursery. All plantations,
except the one at Rosengarten, were
established on large clearings. The
Chapter 4. Provenance Trials 87
five Washington provenances represented one of these
overstory and establishment of natural elevational spans: <100 m, 100-300 m, 300-600 m,
regeneration
600-900 m, 900-1,500 m. The seed stemmed from
of pine affected the growth of Douglas-fir.
In spite of all these shortcomings, a broad pic-
ture of differences in provenance performance did
emerge. Coastal provenances from low elevations
in Washington grew best in the coastal region of
northwestern Germany and at low elevations in
the montane regions of central Germany.
Provenances from interior southwestern British
Columbia showed their best performance in the
northeast German plains and at high elevation
sites in central and south German mountain
ranges.

The 1930 Geyr von Schweppenburg trial


The test was established with 13 provenances in
a single plantation at an elevation of 360 m in the
Gahrenberg forest district in Lower Saxony by
Geyr von Schweppenburg, Professor at Georg-
August University at Göttingen. Five of the
provenances belonged to var. menziesii, four to
the northern sub- group, and three to the southern
subgroup of the var. glauca. Except for two of the
inland provenances obtained from the Danish seed
firm Rafn, all other seed came from the Long Bell
Lumber Company and probably from seed
sources geographically close to those used in the
Wiedemann trials. The 1953 mea- surements at
Gahrenberg by Schober (1954) gave results
similar to those reported from the Braunlage and
Freienwalde plantations—that is, the highest
volume production by the Salmon Arm,
Kamloops, and low- and medium-elevation
Washington prov- enances, and marginal growth
by the Oregon and high-elevation Washington
provenances.

The 1930 Fabricius trial


Professor Ludwig Fabricius, former Director of
the Bavarian Forest Research Institute, established
a provenance trial with test plantations in 23
Bavarian forest districts (Rohmeder 1954). The
provenances in- cluded five from the western slope
of the Washington Cascades, one from the
Siskiyou Mountains, and one from Mexico. The
Washington provenances were not identified by
geographic location but only by a letter code,
based on the five elevational spans. Each of the
Siskiyou
a 1929 shipment by the Longbell Lumber Provenance a e i o u Mountains Mexico
Company. The Longbell Lumber Company 3
MAI m /ha 4.7 7.3 4.3 3.2 2.5 1.7 1.0
used letter codes to indicate elevation of seed
source and the correspond- ing number of
frost-free days (Table 4.3).
The objective of the trial was to learn more
about potential seed sources specifically for
use in Bavaria. Presumably, limited
availability of planting stock permitted
planting of all 7 provenances solely at the
Grafrath test plantation. Plantations in the
other 22 districts contained just one or two of
the provenances. At the time of measurements
in 1951, test plantations were left in only 12
of the districts. Performance in regard to total
volume production at age 21 from seed is
shown in Table 4.4.
Phaeocryptopus gaeumannii, the Swiss
needle cast pathogen, was present in all
plantations, but prov- enance e was less
affected by the disease than any of the others.
That circumstance may have been a
contributory factor to provenance e’s superior
vol- ume production. In spite of its
shortcomings, the Fabricius trial pointed to
seed from low and medium elevations in
western Washington as a good choice for
planting on medium elevation sites in Bavaria.
The 1954/1958 Lower Saxony Forest
Experiment Station trials
Renewed interest after World War II in
finding suitable seed sources of Douglas-fir
for planting in Germany led to the initiation of
a provenance trial in four Lower Saxon
forest districts in 1954

Table 4.3 Letter codes used by Longbell Lumber


Company to indicate elevation of seed source and the
corresponding number of frost-free days.

Code Elevation (m) Frost-free days


a <100 <270
e 100-300 180-270
i 300-600 150-180
o 600-900 110-150
u 900-1,500 60-110
y >1,500 >60

Table 4.4 Performance of seed in test plantations in 23


Bavarian forest districts in regard to total volume
production at age 21.
88 Douglas-fir: The Genus Pseudotsuga
through the initiative of Ernst Pein, owner of a commercial
large forest-tree nursery near Hamburg. The trial
included 19 provenances from Washington and
Oregon. A complement to that trial consisted of
establishing test plantations in eight more forest
districts by the Lower Saxon Forest Experiment
Station in 1958 with four Washington, one
Oregon and two British Columbia provenances.
Seed for the trials came from commercial sources.
In both the 1954 and 1958 tests, the
Washington provenances grew better than those
from Oregon, and low-elevation provenances
grew better than those from higher elevations. Of
the two British Columbia provenances, Salmon
Arm grew better than the Vancouver Island
provenance (Dong 1970). Mortality and severity
of frost damage was sig- nificantly less in the
Washington provenances than in those from
Oregon. Lowest mortality and frost damage
occurred in Washington provenances from
elevations below 600 m and in the Salmon Arm
provenance. The southern Oregon provenances
incurred the heaviest losses (Dong 1973).
The 1955/1958 Baden-Wuerttemberg trials
Aside from the provenance plantation at
Kirchzarten—established in 1932 on an extreme
site—no other existed in Baden-Wuerttemberg be-
fore 1955. This region has some excellent stands
of Douglas-fir, but they can provide only a
fraction of the amount of needed seed. The
initiation of 4 provenance tests on 16 sites
throughout Baden- Wuerttemberg from 1955 to
1961 represents an effort to provide some basis
for selecting suitable seed sources (Kenk and
Thren 1984a).
The 1955 test established by Professor
Mitscherlich of Ludwig Albert University Freiburg
and Mr. Kirschner, a retired forest service officer,
included 9 Washington and 10 Oregon
provenances. The seed stemmed from collections
by the USDA Forest Service and was obtained by
Ernst Pein of the Pein & Pein Forest Tree
Nurseries during a visit to the Pacific Northwest.
Information about site and quality of mother
stands was provided by Charles Rindt, former
silviculturist of Region 6 of the USDA Forest
Service.
The second 1955 trial, with two British
Columbia and three Washington provenances with
seed of loosely defined origin, was established by Because of spotty cone crops, seed collection took
the Stuttgart Forest Directorate. Professor from 1955 to 1958. Trees were outplanted as 1+2s
Mitscherlich initiated a third test in 1958 with four in 1961. Measurements were taken at ages 11, 16,
provenances from Washington and one from and 21 from seed.
Oregon. The seed purchased from the Manning
Seed Company in Washington was a composite
from several stands.
A fourth trial was established 1961 at 10
locations in Baden-Wuerttemberg through the
initiative of Prof. Schober (Kenk and Thren
1984a). The 11 North American provenances,
4 from British Columbia, 4 from Washington,
and 3 from Oregon in this test stemmed from
seed collections made for the 1958
provenance trial in northwest Germany
(Schober et al. 1983). Progeny from 4
Douglas-fir stands in the Black Forest and the
Suabian Alb was planted, in addition to the
North American provenances.
Kenk and Thren (1984b) concluded that
after 22 years of observation of test
plantations in the 16 localities preliminary
judgments were possible as to the suitability
of seed sources for cultivation in Baden-
Wuerttemberg. Washington provenances from
elevations between 300 and 600 m, as well as
progeny from German stands, appear to be
best suited because they performed extremely
well on all test sites. Poorly suited seed
sources are from interior British Columbia,
the Oregon Cascades, and the Klamath
Mountains because of their unsatisfac- tory
growth.
The 1958 Schober trial
Members of the Section Mensuration in the
German Union of Forest Research Institutes
decided in 1954 to initiate a Douglas-fir
provenance trial because of the need for more
extensive testing of provenances of ex- actly
known origin. The task fell to Reinhard
Schober, Professor at Georg-August
University Göttingen, who, in conjunction
with the Lower-Saxon Forest Experiment
Station, organized the establishment of an
experiment with 39 provenances on 14 sites in
northern and western Germany (Schober et al.
1983, 1984).
The trial includes seed from 37
documented sources in British Columbia,
Washington, Oregon, and two German stands.
Chapter 4. Provenance Trials 89
the
The results confirmed those of earlier trials but
revealed something less apparent in earlier tests.
The provenances from Vancouver Island, and
especially those from interior British Columbia,
decreased growth with increasing age. Secondly,
provenances from the coastal region of
northwestern Oregon showed surprisingly good
growth, equal to that of those from the
Washington Cascades and the Olympic Peninsula.
Differences in growth between provenances were
considerable, although the trial covered only that
part of the species’ natural range of interest for
procuring seed to be used in Germany.
Provenances ranged from site class I to III in total
volume production.

The 1958 Hessian Forest Experiment Station trial


The Hessian Forest Experiment Station initiated a
Douglas-fir provenance test in 1958, with 30 of
the 39 provenances used in the 1958 Schober trial.
The material used in the Hessian test stemmed
from 5 British Columbia, 17 Washington, 7
Oregon, and 6 additional German seed sources, 5
of which came from the Odenwald and one from
the Vogelsberg area (Rau 1985). Seedlings were
outplanted in 1961 as 1-2s in seven Hessian forest
districts. Height, diameter, and stem form were
recorded at age 23 from seed.
The best provenance for growth was
Humptulips from the Olympic Peninsula. Most
provenances from the west slope of the
Washington Cascades, especially those from the
Darrington area, likewise showed excellent
growth. The performance of the two interior
provenances from British Columbia was poor, and
growth of those from Vancouver Island and the
western half of the Oregon Cascades was only
slightly better. Progeny of German stands,
particularly those from the Odenwald, was gen-
erally superior in growth to nearly all the North
American provenances. Results of the Hessian
test generally agreed with those of the 1958
Schober trial, in spite of differences in
experimental design and site conditions.

The 1962–1963 trial of the Schmalenbeck Institute


of Forest Genetics and Forest Tree Breeding
Klaus Stern, Professor of Forest Genetics at
Georg-
August University, Göttingen, initiated a trial in
the Olympic Mountains, and the Oregon Cascades
Forest District Nordhorn as part of the were judged to contain less desirable seed
Emsland re- forestation project in northwest sources. Seed sources in the Shuswap Lake area
Germany (Herrmann 1973). The test included and the west
81 provenances from throughout the range of
the species including one from Mexico at lat
25°17 N. The seed stemmed from collections
made by North American forest research
institutions.
Height measurements at age 10 and 11
years showed the provenances from the
coastal regions of Washington and northern
Oregon to be the best performers (Stern et al.
1974, Hattemer and König 1975). The growth
of the provenances belonging to the interior
variety was moderate or poor. Three
provenances from New Mexico which ranked
right behind the best Washington and Oregon
provenanc- es made up a notable exception.
Perhaps mention ought to be made here that
the only New Mexico provenance in the
Schwappach 1910 trial showed far better
growth than any of the other interior prov-
enances in that test until infection with
Rhabdocline pseudotsugae in 1930.
The 1961 German Democratic Republic trial
The Forest Sciences Institute at Eberswalde
estab- lished six plantations in 1961 with 1+2
seedlings across the former German
Democratic Republic from the Baltic Sea
coast to the mountains in the south of the
country (Dittmar et al. 1985). The experiment
included 26 provenances, 2 from interior
British Columbia (Salmon Arm area), 4 from
the east side of Vancouver Island, 14 from
Washington, and 6 from Oregon.
The experiment demonstrated relatively
small dif- ferences between provenances at
plantation age 20. That finding was a surprise
because the provenances represented only that
part of the species’ natural range shown in
earlier provenance tests to contain the optimal
seed sources for its cultivation in central
Europe. The results, however, narrowed the
choice of seed sources apparently best suited
for the north- east German lowlands, referred
to as “Pleistozän”, to the west slope of the
Washington Cascades, the Washington Coast
Range, and the Cascade Range of northwest
Oregon. Vancouver Island, the north side of
90 Douglas-fir: The Genus Pseudotsuga
slope of the Washington Cascades were Germany at elevations ranging from 20 to 600 m,
considered to be the most appropriate for the were established with seedlings from the 1970
mountainous region in the south of the country. sow- ing. Their performance in 7 of the 16
plantations was reported at age 9 from seed
The IUFRO Trials (Kleinschmit et al. 1979). Provenances displayed
As noted, the provenance tests of the first seven distinct differences in their ability to adapt to site
de- cades of the 20th century had covered the vast conditions at the 7 plantations. Provenances from
range of Douglas-fir in a rather spotty fashion; the coastal region and the North Cascades of
thus, their results had provided a limited basis for Washington displayed the greatest adaptability, if
selecting seed sources suitable for use in the pre- it is defined as supe- rior growth combined with a
1989 Federal Republic of Germany. In addition, high rate of survival. Those from Vancouver
insufficient or even complete lack of information Island, the central Cascades in Washington, and
about the exact geographic origin of many the coastal region of north- ern Oregon ranked
provenances in the tests diminished the utility of next; they had good growth but a lower rate of
their results. To alleviate these shortcomings, the survival. Provenances from coastal British
tree breeding institutes of the states of Baden- Columbia, the southern Washington Cascades,
Württemberg, Bavaria, Hesse, and Lower Saxony and the Cascades in Oregon adapted poorly to the
decided to participate in the IUFRO international test sites. Their growth was very un- even and
Douglas-fir provenance study. The four tree- their rate of mortality was extremely high. The
breeding institutes planned a joint experiment unsatisfactory performance of the provenances
with 111 provenances from the IUFRO from the southern Washington Cascades
collections, 9 non-IUFRO provenances from contrasted with the findings of a study by Racz
British Columbia, and seed from 4 German stands and Kleinschmit (1978), which included 18
(Kleinschmit et al. 1974). Seed from the IUFRO provenances from that region with satisfactory
collections consisted of 38 provenances from the performance. Provenances from Oregon south of
northern subgroup of the inland variety and 72 lat 45° N and California were judged to be
provenances of the coastal variety, and one unsuitable for cultivation in Germany because of
Mexican provenance whose taxo- nomic status poor growth and a high rate of mortality.
within the genus Pseudotsuga is still open to Provenances belonging to the interior variety had
question. Provenances from the southern a very low rate of mortality, but their growth was
subgroup of the inland variety were not included too low to recommend them for use in Germany.
in the experiment because results of older trials Performance of provenances in six of the plan-
(Kleinschmit 1973) had indicated their poor suit- tations at age 14 (Kleinschmit et al. 1987) and age
ability for cultivation in Germany. 20 (Kleinschmit et al. 1990) followed essentially
The experiment was divided into two parts. For the pattern observed at age 9. Provenances from
part one, all 124 provenances were sown 1970 in the western part of the Olympic Peninsula and the
each of three nurseries located in northwest, western slope of the northern Washington
southwest, and southeast Germany. The objective Cascades continued to display the best overall
of part one was producing seedlings for performance, and those from the Puget Sound
establishing planta- tions to be observed for 20 region, southern Oregon, and California the
years, and to make a first selection at the end of poorest.
the nursery phase. Part two of the experiment Seedlings representing 50 provenances selected
consisted of a sowing in 1973 of the best one-third for part two of the German participation in the in-
provenances from the 1970 sow- ing selected on ternational IUFRO provenance test were
the basis of height growth and frost hardiness. outplanted in 1975 in 12 plantations (Kleinschmit
Performance of seedlings raised for part two of 1978). Results have not yet been published.
the experiment was supposed to be followed for
The 1970 Hesse trial
40 years in the field (Kleinschmit et al. 1974).
In spring of 1973, 16 plantations, distributed Aside from its participation in the IUFRO prove-
throughout the pre-1989 Federal Republic of nance test jointly by the four German forest
breeding
Chapter 4. Provenance Trials 91
600 m elevation
institutes, the Hessian Forest Breeding Institute
initi- ated an additional provenance experiment in
1970, designed to narrow the choice of seed
sources for use in Hesse (Jestaedt 1980). The
experiment includes 118 provenances, 25 belong
to the northern subgroup of the interior variety and
93 to the coastal variety from the IUFRO
collection, 9 provenances from com- mercial
sources, and progeny from 4 German stands in
Baden, the Palatinate Forest (Pfälzer Wald), the
Eifel mountains, and Lower Saxony. Added to the
experiment was progeny of 91 single-tree
selections from 6 Douglas-fir stands in the Eifel
Mountains. At the time of cone collection in 1968,
the 6 stands ranged in age from 32 to 86 years.
As in the col- lections made by IUFRO in 1966-
1968, cones were collected from 8 to 20 trees in
each of the 6 stands (Rau 1987). Test plantations
were established in 1973 with 1+2 seedlings in 13
locations throughout Hesse. Measurements of
several quantitative and quali- tative traits during
the nursery phase and in the field provided, as
early as age 8, useful criteria for judging
provenances (Jestaedt 1980). Provenances from
the lower elevations of the western slope of the
North-Washington Cascades and the Olympic
Peninsula were by far the best performers. Next in
growth ranked the provenances from the east side
of Vancouver Island, the lower elevations of the
west side of the Cascades in southern Washington,
the Coast range in northwest Oregon, and progeny
from the 4 German stands. Provenances that
belong to the inland variety and coastal
provenances from the southern part of their
range had the poorest growth. Measurements at
age 14 did not significantly
change rankings (Rau 1987).
Results of the 1970 trial largely agreed with
those from the 1958 provenance test in Hesse,
which led Jestaedt (1980) to recommend use in
Hesse of seed from the following regions:
• the area between the coast and west slope of
the Cascades in northern Washington and
southwest British Columbia between the
Frazer River and lat 47° N, up to about 700
m elevation
• the southern and southwestern part of the
Olympic Peninsula, to about 700 m
elevation
• the east side of Vancouver Island, to about
seed sources in the range of the northwestern
Height and diameter growth, phenological subgroup of var. glauca appear to be better
char- acteristics, and frost resistance of suited than any var. menziesii seed sources. Var.
progeny from the single-tree selections in the menziesii progeny from German Douglas-fir
6 Eifel Mountain stands were recorded at stands performed in provenance
ages 8 (Jestaedt 1980), 14 (Rau
1987), and 20 (Hesse FVA 1993). Growth of
these trees matched that of the best IUFRO
provenances. Although marked differences in
the measured traits were found between
progeny from individual trees of each of the
six stands, mean values for progeny from each
of the stands varied little. Jestaedt (1980)
stated that the range in age of the six mother
stands justifies the assumption that they come
from differ- ent seed sources and that the
homogeneity of their progeny represents a
process of adaptation towards a “land race,”
as defined by de Vecchi (1969).
The GDR IUFRO Trial
Another German IUFRO provenance test was
initiat- ed in 1970 in the former German
Democratic Republic by the Forest Sciences
Institute at Eberswalde (Braun 1985). The
experiment included 139 IUFRO prov-
enances and progeny of 5 East German
Douglas- fir stands. Seedlings were
outplanted in 1973 near Stralsund in the
northeast German low lands. The objective of
the experiment was to test the ability of these
provenances to grow in open areas without the
benefit of any kind of shelter.
Height and survival was recorded at age 13
from seed. The best performers were
provenances from northwestern Washington
because of good growth coupled and an
acceptable rate of survival. Provenances from
interior British Columbia had the highest rate
of survival (77-90%) but only moderate
growth. Heavy losses, primarily caused by
frost, accounted for the low rate of survival
(34%).
Seed sources for Germany
The German Douglas-fir provenance trials
demon- strated that the best seed sources for
that country are to be found in northwestern
Oregon, western Washington, and
southwestern British Columbia at elevations
up to about 600 m. For high elevation sites in
German mountain ranges, however, some
92 Douglas-fir: The Genus Pseudotsuga
tests as well, or even better, than the best North The
American var. menziesii provenances. Unfortunately,
seed production from German Douglas-fir stands
is insufficient to meet demand, and probably will
remain so for years. Therefore, the former GDR
will continue to depend on seed imports for the
foresee- able future.

International Studies
British provenance trials
Douglas-fir has grown in Britain since 1828, but
not until 1928, a century later, did the provenance
tri- als begin. The next 44 years saw the
establishment of 11 experiments with 82
provenances in Scotland and northern England,
and of 19 experiments with 118 provenances in
southern England and Wales (Lines 1987).
Among these are two main series, one planted in
1953/54 and another in 1970-1972. The 1953/54
series constituted the first large-scale British
provenance trial. The seed was obtained from the
Manning Seed Company and consisted of 14
Washington and 3 Oregon provenances. The
establishment of two plantations at Laiken (Nairn)
and Glentress (Peebles) in 1953 was followed, in
1954, by a third planting at Sunart (Argyll) to in-
clude a site with the typical high rainfall and mild
winter of western Scotland (Lines 1957). The
1953 planting at Glentress also included two
provenances from interior British Columbia
(Salmon Arm and Prince George) and three from
Vancouver Island (Lines 1956).
In 1954, plantations were established at three
sites
in England (Mortimer, Herfordshire; St. Clement,
Cornwall; Shouldham, Norfolk) and on a site in
Wales (Rheidol, Cardiganshire). A first
assessment at the end of the third growing season
showed a considerable difference in growth
between the Washington and Oregon
provenances, the Oregon provenances being
rather poor (Wood et al. 1960). At plantation age
12, the Washington provenances had maintained
their superiority over the Oregon provenances at
the four English and Welsh sites. As a follow up
to the 1954 trial and to extend it to a wider range
of seed sources, 15 provenances were planted in
March 1968 at the New Forest and the Forest of
Dean besides a smaller trial at Helwill, Cornwall.
provenances included two from Vancouver 4. General yield class (GYC) is defined as the maximum mean annual
increment per hectare in m3.
Island, four from Washington coastal areas, 5. Abandoned after 6 years.
and nine from along the Oregon coast to its
southern end (Lines and Mitchell 1968).
Measurements of the 1953 (age 20) and 1954 (age
23) showed that differences in height between
them were small but total volume differed
considerably between provenances, with Elma
outstanding at 314 m3/ha = GYC 20 (Lines
1977).4
The second main series consists of
plantations established in 1970-1972 with
seed from the 1966 and 1968 IUFRO
collections. The intent was to sample portions
of the species’ range not covered by earlier
trials and to test different provenances from
known regions of promise.
In 1970, three plantations were established
in Scotland, one at Culloden Forest 32 km
east of Inverness and two at Craigvinean
(Dunkeld Forest) 24 km northwest of Perth.
Each site contained six provenances from
British Columbia, eight from Washington and
six from Oregon. Seven more IUFRO
provenances, three from British Columbia and
four from Washington, together with two
com- mercial seed lots from Elma were
planted in 1972 at Craigvinean (Lines and
Samuel 1987).
Survival and growth of provenances at
Dunkeld Forest differed greatly 6 years after
planting in the field. Little evidence was seen
of clinal variation with latitude, elevation, or
distance from the coast. The trial at Culloden
Forest had to be abandoned because of severe
losses from frost damage (Lines 1980).
The companion plantations in southwest
England were established in 1972 at Bodmin
Forest, Cornwall; Charmouth Forest, Devon;
Forest of Dean, Gloucester; Quantock Forest,
Somerset;5 and Radnor Forest, Hereford. At
Bodmin Forest, seedlings were planted on
east; south- and north-facing slopes, which
closely matched in physical and chemical soil
properties. The provenances included in the
English trials consisted of 44 IUFRO
provenances from interior (8) and coastal (7)
British Columbia, Washington (12), Oregon
(9), and California (8), besides 2 commercial
seed lots from Washington
Chapter 4. Provenance Trials 93
have been modified and reduced to 10 (Figure 4.4).
(Elms and Hoodsport). Not all provenances were That
present on all sites, however, because only 31
prov- enances were planted on each site. At the
end of the sixth growing season in the field,
growth had varied between sites and aspects at
Bodmin Forest. The rela- tive performance
between provenances was constant across the
range of sites, and aspects, however. The best
provenances were from a geographically nar- row
area of coastal Washington and coastal northern
Oregon that has been traditionally used as a
source of seed for use in Britain (Pearce 1980).
A report of height and diameter growth 10
years after planting at four English (Bodmin,
Charmouth, Dean, Radnor) and one Scottish site
(Craigvinean) showed little change in rank of
height growth from that measured after 6 years
since planting (Lines and Samuel 1987). Rank for
diameter was closely correlated with rank for
height.
Provenances from above lat 50° N and below
lat 43° N, as well as high elevations, displayed a
pattern of poor growth. Performance of
provenances from Vancouver Island was only
moderate. Best growth was mainly by
provenances from sites under 305 m elevation in a
U-shaped zone from Arlington southward along
the west side of the Washington Cascades to the
Columbia River and thence north to Forks on the
Olympic Peninsula. Lines and Samuel (1987)
pointed out that results of the IUFRO ex-
periments indicate negligible interaction between
provenances and site and concluded that selecting
different seed sources for different parts of Britain
is likely to be of minor importance.
Height, diameter, and branching characteristics
(branch number, diameter, angle) were measured
on 38 IUFRO provenances 16 years after planting
at the Forest of Dean and Radnor Forest. In
addition, measurements of dbh and basal area
were recorded at Craigvinean 19 years after
planting (Fletcher and Samuel 1990). The trials at
Dean and Radnor contain 38 of the 44 seed
sources with 26 of these common to both sites.
Craigvinean contains 20 provenances 15 of which
are present at either one or both of the Dean and
Radnor sites.
In an earlier report (Lines and Samuel 1987),
the 44 IUFRO seed lots were grouped in 12 zones
based on geographic or climatic data; these zones
change appears to be justified because, for all
the traits measured, significant differences
between zone means were found but rarely
between provenances within zones.
The provenances from British Columbia
(Zone I and II) were the poorest because of
small diameter growth and low basal areas.
They had superior branching characteristics
and high survival but that

1004 I
1006 1109
II 1012 1005
1023
1029
50° 1031 1013
1014
1027
1030
1043
British Columbia
1051 1050
1062 1054 1053 III
IV 1056
1063

VI 1086 1075
1081 V
1089 1083
1094 Washington
45°
Oc 1098
VII 1100
1118
1101

VIII
1130
1124 Oregon
1135
1104
IX 1126
X1144
40° 1147 1146

1148

San Francisco

California

35°

Los Angeles

160 0 160 320 K

Figure 4.4 IUFRO seed lots used in British provenance


trials grouped into 10 zones (from Lines and Samuel
1987).
94 Douglas-fir: The Genus Pseudotsuga
would not offset the low volume production. By species accounted for 32% of the annual planting
contrast, provenances from Washington (Zones program (O’Driscoll 1973), is a likely reason that
III, IV, V, and VI) had high survival and, provenance trials with Douglas-fir did not
combined with above-average diameter, did commence until the early 1960s. The first two
produce the highest basal area. Although the trials were made with commercially collected
provenances from south- ern Oregon and northern seed. Results were of limited usefulness, however,
California had the largest individual tree because of lack of au- thenticity of seed sources.
diameters, the nature of their branch Availability of the 1966/68 IUFRO seed collection
characteristics, a combination of high branch provided the Irish Forest and Wildlife Service
num- bers and large branch diameters, negated the with the opportunity to begin an experiment with
benefits of rapid diameter growth. seed of well-documented origin.
Results of the measurements at age 16 of The trials with seed from commercial
progeny from zones III, IV, and V at these three collections had shown that interior provenances
sites con- firmed predictions from earlier reports were unsuit- able. Emphasis was therefore placed
that the best sources of seed for use in Britain are on selecting var. menziesii provenances with
to be found in a U-shaped area in Washington; to preference to those from low elevations. The 32
that can now be added the southwestern corner of seed lots from the 1966 IUFRO collection,
Washington and the adjoining northwestern corner received in l967, represented 13 provenances from
of Oregon. The areas considered to contain the British Columbia, 11 from Washington, and 8
best seed sources for Britain correspond to the from Oregon. Twenty-five of the provenances
Northwest Forest Tree Seed Council6 zones 202, belonged to the var. menziesii and 7 to the
403, 411, 412, 030 (northern northern subgroup of var. glauca.
half), 012, 041, and 052 at elevations below 500
Seeds were sown in 1968 and outplanted as
m in all these zones (Fletcher and Samuel 1990).
1+1+1 transplants in spring of 1971 on 5 different
A third series of noteworthy experiments was
sites throughout Ireland (O’Driscoll 1978). At the
established to compare progeny from eight old-
year-3 height assessment in the nursery, an
er stands of Douglas-fir, many of which derived
emerging trend showed that high elevation and
from the David Douglas introduction to
more northerly prov- enances occupied lower
Lynedoch, Scone Estate, with five North
rankings than those from low elevations and more
American prove- nances. Seedlings were planted
southerly origin. The latter two kinds of
in 1959 at Elibank (Peebles-shire), Castle O’er
provenances had a higher incidence of lammas
(Dumfries-shire), and Thornthwaite (Cumberland)
growth than the former two kinds.
and in 1960 at the Forest of Dean (Gloucester).
A sixth set of the 32 provenances was planted
An assessment of height growth after six growing
on the grounds of a former nursery at Glenealy,
seasons indicated a rather disappointing
County Wicklow, to study the growth patterns of
performance of progeny from some of the eight
Scottish stands. Many trees had poor growth and the differ- ent provenances. Ranking of
were malformed (Lines and Mitchell 1967, Lines provenances by height growth did not follow a
et al. 1967). Progeny from the Scottish mother clinal pattern. Three prov- enances, Coquille in
trees apparently show signs of inbreeding southern Oregon’s Coast Range, Granite Falls,
depression because of their extremely small and Sedro-Woolley in Washington’s Puget Sound
genetic base. Seed had been collected from only region were consistently in the top three ranks
one or two trees. Progeny from Scottish stands while the interior provenances were always at the
known to result from a large seed import did not bottom of the rankings. Provenances from
show these features (Lines 1987). southern Oregon were the best performers but
their long period of growth increases their chance
Irish provenance trials of injury by early frost. Provenances from the
The decline in the amount of Douglas-fir planted Puget Sound area appear to be more suitable for
in Ireland, after the period 1922 to 1927 when the Irish Conditions as they combine vigorous growth
with a shorter growth period (O’Driscoll 1978).
6. Later the Western Forest Tree Seed Council.
Provenances from coastal Washington,
coastal
Oregon, and the Puget Sound region were the
best
Chapter 4. Provenance Trials 95
and were continued for 12 years at 3-year intervals (de
in the IUFRO provenance trials 9 years after Vries 1961) The MAI30 for the three best prov-
plant- ing (Pfeifer 1988). He considered enances (Chilliwack, BC; middle WA, a composite
provenances from the north and central
Washington coast (seed zone 012, 030) and the
south Washington coast (seed zone
041) as preferable for Ireland, based on the results
of the IUFRO trial.

Dutch provenance trials


Douglas-fir has been planted in the Netherlands
since 1860 with varying success. The importance
of seed source for the success of plantations was
recognized in the beginning years of the 20th cen-
tury. Because origin of the early seed imports was
unknown, a series of provenance tests was begun
in 1923, with 4 interior and 2 coastal provenances.
Every year from 1925 to 1932, a few provenances
were added to bring their total to 35. They
consisted of 11 inland provenances: 10 from
British Columbia and 1 from Washington; and 22
coastal provenances:
2 from British Columbia, 17 from Washington, 2
from Oregon, and 1 from California. Two of the
35 provenances came from Dutch Douglas-fir
stands. Seeds were obtained from several
providers, the seed dealer Katzenstein and Co., in
Atlanta, Georgia; the “Associated Foresters” in
Calgary, Alberta, the Long Bell Lumber Company
in Washington, and the USDA Forest Experiment
Station in Portland, Oregon. The provenances
were used in establishing 27 test plantations in
five Dutch provinces. Veen (1951), in his analysis
of the performance of the 35 provenances,
distinguished four groups—very good, good,
medium, poor —based on height growth between
10 and 15 years of age. The best perform- ers
were Washington provenances from low eleva-
tions, and the worst were those from interior
British Columbia and northeastern Washington.
Veen (1951) concluded that inland and high-
elevation coastal provenances are poorly suited
for planting in the Netherlands. Both grow slowly,
the inland prov- enances are highly susceptible to
Rhabdocline needle cast, and coastal provenances
from high elevations are highly prone to injury
from late spring frosts because of their early
flushing.
Thinnings in the test plantations began at age
20
of King, Lewis, Thurston, and Pierce to then came from Pacific County. That
Counties; and Pacific Coast, WA) ranged from provenance “was tested in the 1923-1932 trial
10.1 to 12.2 m3/ha. The MAI30 for the two series and is known under the name Pacific
poorest provenances (north- east Washington Coast.”
—composite of Okanagan and Ferry
Counties); and Salmon Arm, British
Columbia) ranged from 3.5 to 5.9 m 3/ha. The
measurements of volume growth provided
essentially the same ranking of provenances
as Veen’s (1951) based on height growth.
Results of the provenance research begun
in 1923 indicated that western Washington
and southwest- ern British Columbia
contained the most promising seed sources for
use in the Netherlands. But the sampling had
not been extensive enough to allow for a
reliable delineation of seed collection areas
for the Netherlands. Achieving that goal
required more provenance research. The
IUFRO seed collections provided a timely
opportunity for testing many provenances
from throughout the species range.
In 1971, a provenance test was begun with
57 provenances from the 1966/67 IUFRO
seed collec- tion (Kriek 1974). Twenty-five of
the provenances came from British Columbia,
24 from Washington, and 8 from Oregon. All
provenances belonged to the coastal variety.
Seed was sown in two nurseries in December
1967, and seedlings were transplanted in
spring 1969. Two plantations were established
with the transplants, one in 1971 at
Sleenerzand (52°50’ N lat.) in northeastern
Holland, the other at Sprielderbos (52°14› N
lat.) in the southwestern part of the country.
Kriek (1974, 1978) reported on the
development of the provenances up to age 10
from seed. Provenances from Washington
performed best, but not equally well,
throughout the areas sampled by IUFRO in
that state. The best provenances came from
the western slope of the northern half of the
Cascades covering parts of seed zones 202,
403, and 412 at altitudes of 100–300 m, from
along the west- ern and southern flanks of the
Olympic Mountains covering parts of seed
zones 012 and 030, and Pacific County in the
southwest corner of the state just north of the
Columbia River, covering part of seed zone
041. Kriek (1974) remarked that the very best
provenance imported into the Netherlands up
96 Douglas-fir: The Genus Pseudotsuga
The measurements in 1987 up to age 20 from 650 m. The seed was
seed closely matched the results up to age 10 both
at Sleenerzand and Sprielderbos. This outcome
jus- tifies, according to de Vries (1990),
continuing to regard provenances from these three
areas as the best sources of seed for the
Netherlands.
In 1970, a second provenance study was initi-
ated with 104 provenances, 58 of which came
from individual trees, from the 1968/69 IUFRO
collection. Twenty-five of the provenances—4
from Oregon and 21 from California—belonged
to the coastal variety. The other provenances—19
from Colorado, 4 from Utah, 28 from Arizona,
and 28 from New Mexico—were from the
southern subgroup of the interior variety. By age
11 from seed, their perfor- mance already
indicated that none of the areas where these
provenances originated should be chosen as seed
sources for use in the Netherlands (Research
Institute for Forestry and Landscape Planning
1981, Kriek 1983).
In the late 1980s, two of the three plantations
es- tablished with progeny from the 1968/1969
IUFRO collection were abandoned because of too
many loss- es of trees. The third plantation has
been maintained as a demonstration object for the
consequences of planting provenances unable to
adapt to sites in the Netherlands (de Vries 1990).
Belgian provenance trials
The first provenance tests in Belgium date to 1925
(Galoux 1956). They included provenances from
the range of interior Douglas-fir and from
throughout the range of coastal Douglas-fir. Their
results indi- cated that interior Douglas-fir is
unsuitable for use in Belgium because of inferior
growth and high susceptibility to Rhabdocline
pseudotsugae. The out- come of these first tests
suggested that progeny from Douglas-fir in
western Washington and south- western British
Columbia held the most promise for cultivation in
Belgium.
To better identify provenances desirable for
dif- ferent parts of Belgium, the Forest Research
Station at Groenendaal initiated, in 1951, a trial
with 31 prov- enances, 22 from Washington and 9
from Oregon (Gathy 1961). The mother stands
stocked at altitudes ranging from about 100 m to
500 m, except for two stands in Oregon at about
provided by two seed dealers, Manning Seed d’Amélioration des Arbres forestiers), a branch of
Co. and Wood Seed Co., in Washington. INRA (Centre National de la Recherche
Two plantations were established in 1954
with 3-year-old plants in the High Ardennes at
460 m and 515 m, respectively. Two more
plantations were established in 1955 in central
Belgium at 65 m and 75 m, respectively.
Washington provenances from elevations
below 200 m showed the best growth and
were also the most frost resistant. None of the
Oregon provenances matched the Washington
provenances in growth or frost resistance.
That pattern had not changed significantly by
age 22 from seed (Nanson 1978). The best
provenances originated from a U-shaped area
extending from Forks near the Pacific coast to
Darrington. Trees from those seed sources
grew best at both low and high elevations in
Belgium; they flushed late and thus were least
susceptible to injury by late frosts.
The Groenendaal Forest Research Station
initiated in 1969 a provenance trial with 26
provenances: 5 from southwest British
Columbia, 6 from Vancouver Island, 12 from
western Washington, 2 from eastern
Washington, and 1 from Oregon, with seed
from the 1967/68 IUFRO collections. The
trial included also progeny from 10 Belgian
Douglas-fir stands in the Ardennes (Nanson
1973).
Performance in the nursery at age 3 from
seed was best by provenances from western
Washington, aver- age for the Belgian
Douglas-fir, and below average for the British
Columbia and Vancouver Island prov-
enances. The poorest provenances were those
from eastern Washington and Oregon. The
Brookings, Oregon, provenance suffered
extreme injury from a fall frost. Nanson
(1978) concluded that elevations below 500 m
in western Washington—roughly cov- ered by
seed zones 030, 240, 232, 412, 411, 403, and
202—are most likely to contain the seed
sources most suitable for use in Belgium.
French provenance trials
The rapid growth of the share of Douglas-fir
in re- forestation after World War II in France
prompted the initiation of provenance trials
with that species in the 1960s. The Institute of
Forest Tree Improvement (Station
Chapter 4. Provenance Trials 97
de Lacaune, and Le Treps in the Massif des Maures at
Agronomique) established Douglas-fir altitudes of 800, 700, and 600 m, respectively (Birot
provenance and progeny tests in 15 locations and Ferrandes 1980). The two plantations established
throughout France from 1965 to 1978 (Figure in 1968 at le
4.5).
The first series of provenance plantations was
established in the years 1965 to 1968. The
locations of these provenance tests are Epinal in
the north- east, Peyrat-le-Chateau and Besséde-
Barade in the southwest, and St. Amans Valtoret,
Sauclières and Le Treps in the south. Seed for
these provenance tests came from commercial
collections of coastal Douglas-fir in North
America except for the test at Sauclières where all
plants represent progeny from Douglas-fir stands
in France.
The 1965 test plantation at Peyrat-le-Château
contains 5 British Columbia, 15 Washington,
and 5 Oregon provenances, as well as progeny
from a 25-year-old French Douglas-fir stand in
the south- ern Rhone region. Lacaze and
Tomassone (1967) analyzed the nursery
performance of these prov- enances. They
concluded that the results pointed to Vancouver
Island, the Olympic Peninsula, and the western
slope of the Washington Cascades as the areas
that contain the most suitable seed sources for
eastern France.
Assessment of performance at age 18 from
seed based on volume production, stem form,
branch an- gle, and knottiness showed most of the
Washington provenances as best performers but
also surpris- ingly poor performance by the
Vancouver Island provenances (Birot and Lanares
1980).
Rozenberg (1993) compared height growth of
12 (2 Vancouver Island, 7 Washington, 3 Oregon)
of the 25 provenances to age 25 from seed. The
compari- son was based on 12 felled trees from
each of the 12 provenances. He created height-age
curves for each of the 144 trees, which indicated
that changes in rank of provenances between age
10 and 25 were mostly minor. The best
provenances, those from Washington, remained
the best, and the poorest, those from Oregon,
remained the poorest.
The sites for the provenance trials aimed at
identi- fying of seed sources suitable for the
Mediterranean region of France are at Sauclières
in the Cevennes, St. Amans Valtoret in the Monts
plantations with most provenances were at
5° 0° 5°
Peyrat-le-Chateau, Amance, and the Forêt
d›Orléans.
GERMANY

BELGIUM

50° 50°
LUXEMBOURG

Paris
Amance

Orléans
Epinal
Nor th
Atlantic Ocean FRANCE
SWITZERLAND
Source Faux-la-Montagne
IUFRO
Commercial French
Peyrat-le-Château
CendrieuxCol des Trois-Sœurs
ITALY
45° 45°
Besséde-Barade Saint-André- lés-Alpes
Saucliéres
Bay of Biscay
St. Amans Valtoret
Le Treps

Marseille Lambert
Félines-Minervois Saint
Lareulle
Giron
Mediter r anean
SPAIN
Sea 0100 km

5° 0° 5°

Figure 4.5 Locations of Douglas-fir provenance and


progeny tests established by INRA in France; map compiled
by Hermann (2012) from several literature sources.

Treps in the Massif des Maures were to test 3


prov- enances from the California Coast
Range, 2 from the Sierra Nevada, 1 from the
extreme southern end of the Cascades, 1 from
Washington, and progeny of a French
Douglas-fir stand from the Beaujolais
Mountains. A severe drought in 1970
showed the California provenances to be far
more drought resis- tant than the Washington
Granite Falls and French provenance. The
survival rate of the 6 California provenances
was 75%, compared to 50% for the latter
two (Birot and Ferrandes 1972). An estimate
of biomass at age 12 from seed by the
summation of height of surviving trees
indicated the superiority of the California
populations. Birot and Ferrandes (1980)
concluded, on the basis of their findings, that
Douglas-fir from California seed sources
shows promise for use in the French
Mediterranean region. A second series of
trials was begun in 1970 with seed from the
IUFRO collections and from several Douglas-
fir stands in France. These trials contain
provenances representative of much of the
species› range. Fifteen plantations with
varying numbers of coastal and interior
provenances had been es- tablished
throughout France by 1977 (Figure 4.5). The
98 Douglas-fir: The Genus Pseudotsuga
Temporarily waterlogged soils at the Forêt and Arizona grew better than those from Utah
d›Orléans and Amance led to such poor survival and Colorado.
of the interior provenances that the experimental The performance of southern Oregon and
plots had to be abandoned. By contrast, the California provenances from the IUFRO
difference was slight in survival between interior collection, at age 13 from seed, tend to support the
and coastal provenances, 91% versus 97%, at conclusion drawn by Birot and Ferrandes (1980)
Peyrat-le-Chateau 6 years after planting. In from an earlier study that such provenances may
general, provenances of variety menziesii from be successfully grown in the French
western Washington outperformed those from Mediterranean area. Although variety glauca may
other parts of the range of coastal Douglas-fir and initially grow well at high eleva- tions in the
the variety glauca provenances (Bastien et al. Mediterranean mountains because of its cold
1980). In tests in the French Mediterranean region hardiness (Bastien et al. 1988), its susceptibility to
with provenances from the IUFRO collections, severe infection by Rhabdocline pseudotsugae will
most be- longed to variety glauca and the most likely offset benefits derived from the cold
remainder to vari- ety menziesii from southern hardiness.
Oregon and California. Results after the first The appearance of Rhabdocline pseudotsugae in
decade after outplanting in the field showed the provenance plantations in the Mediterranean
higher survival and better growth by coastal than region led to a detailed study of the pattern of
interior Douglas-fir, except on the most severe infection with the fungus at the St. André-les-
sites, namely Felines Minervois and St. André-les- Alpes test site, which contains 10 variety
Alpes (Bastien et al.1988). Incidentally, St. André- menziesii and 66 variety glauca provenances
les-Alpes is the only site where survival in (Soutrenon 1986). Data collected each spring from
provenances from the southern subgroup of 1984 to 1986, the 9th, 10th, and 11th years after
variety glauca was higher than in those from planting in the field, demonstrated that the
the north- ern subgroup: that is, 72% versus 51%. ranking of provenances in percentage of trees
Within the southern subgroup, provenances from infected remained nearly the same from one
New Mexico

year to the next. The percentage of infected trees
5° 0°
was low in coastal Douglas-fir and the
GERMANY northern subgroup of the inland variety but
BELGIUM
high in its southern subgroup. The exception
LUXEMBOURG
50° was Douglas- fir from Mexico, which
4 apparently was not very susceptible to the
Perruel disease.
Paris
7
Kerpert The AFOCEL provenance tests
Nor th
FRANCE
A French timber industry group, the
Atlantic Ocean Mantes 5
St. Brisson Association Forêt-Cellulose (AFOCEL),
SWITZERLAND
began a program of provenance tests with the
Limoges
2 procurement of 186 seed lots, including 75
1 Lyon
MazerollasRoyere from the IUFRO collec- tions, which covered
ITALY
45°
Bordeaux
nearly the entire natu- ral range of Douglas-
Bay of Biscay fir (Michaud 1978). The AFOCEL established
Arudy
a first series of test planta- tions in various
Toulouse
6 parts of France (Figure 4.6) from 1977 to
3
Bouisse M e d i t e r r a n e a n 1978. From 1978 to 1981, two additional
SPAIN Sea 0100 km
series of tests were initiated, one with 82
5° 0° 5° prov- enances from Washington and the
other with seed collected from French
Douglas-fir stands. A report on performance
of provenances in
Figure 4.6 AFOCEL provenance test plantations established by seven of the plantations of the first series, mea-
AFOCEL in France (from Michaud 1987).
sured by total height at age 8 after planting in
Chapter 4. Provenance Trials 99

the field, indicated that Bay of Biscay


the most vigorous FR
SM FRANCE
RP 5 LG
prov- enances came 43°
6 78
1 RE GA VA CO
from the part of the 2
AT
LV 9 SL OC
species range between CD 11
lat 44° N and 50° N Pontevedra LH SPAIN 10
366 12 13 20
west of the crest of the 4 18
42° 66 BA SE
Cascades (Michaud
Atlantic Ocean

1987). Ranking of 14
PORTUGAL 15 17 19
prov- enances remained
41° 16
nearly the same for all 0 30 60 90 mi 6° 4° 2°
test sites.
lateness As in was
of flushing the 48 100 144 km
INRA trials,

positively correlated Autonomy/Province Map Autonomy/Province Map Plantation Elev. (m)


with growth. no. no.
Principal tests
Recommendations based Galicia Castilla y León
La Coruña 1 León 11 LH La Hermida 700
on provenance tests Lugo 2 Palencia 12 Satellite testsdel Eje
SE Sierra 1360
Pontevedra 3 Burgos 13

Although provenance Ourense 4 Zamora 14 BA Bande 900


Asturias 5 Valladolid 15 CD Castro Dozón 750
tests of Douglas-fir did Cantabria 6 Segovia 16 RP Rocha da 520
not begin until the mid- Soria 17 Perdiz FR 580
Euskadi/País Vasco
La Rioja 18 Fragavella 450
1960s in France, Bizkaia/Vizcaya 7
RE Regavella 850
Araba/Álava 8
recom- mendations for Gipuzkoa/Guipúzcoa 9 SM Sierra de Meira 660
Zaragoza 19 GA Gamalleira 920
seed sources to be used Navarra 10 Huesca 20 VA Valdemadeiro 850
in that country were CO Conforcal 600
already issued 30 LG La Gallina 1160
LV La Vecilla 900
years later based on the SL Salinas de 750
findings Léniz AT Ataun 900
OC Ochagavía
from these trials. The Figure 4.7 Location of Spanish Douglas-fir provenance tests (from Fernandez et al. 1993).
recommendations are
contained in a publication of the French Ministry The first provenance tests in Spain were installed in
of Agriculture (Michaud 1997). They are for six 1950 by Dr. Fernando Molina and served as the basis
regions by listing seed zones, established by the for the common-garden tests initiated in 1976 (Vega
Northwest Forest Seed Council, which are 1990). Of the provenances used in this test, 85 were
considered to contain seed sources suitable for from the 1966/68 IUFRO seed collections;
each of the regions. For the northwest and
northeast of France, the Massif Central, and the
foothills of the Pyrenees, all the rec- ommended
seed sources are in western Washington and
northwest Oregon at elevations below 450 m.
Only in the Mediterranean region are seed sources
from California below 1,200 m and from southern
Oregon below 450 m recommended.

Spanish provenance trials


4 from a 1970 collection by the Institute of
Forest Genetics (IFG) at Placerville,
California; and 1 each from a 1961 and 1965
collection in the Sierra Madre Oriental of
eastern Mexico (Fernandez et al. 1993). Both
coastal and interior seed sources were
included in these tests. Sixteen test
plantations, 10 in north- west and 6 in north-
central Spain, were established from May
1978 to April 1981 (Figure 4.7). The plan-
tations at Carballa Blanca in the Sierra del Eje
in the northeast and La Hermida in the
northwest of Ourense province received all
the IUFRO and IFG provenances obtained for
the 1976 common-garden test. The two
plantations are situated on sites of contrasting
climates. Carballa Blanca has an inland
climate with a short growing season and cold
win- ters, but the climate of La Hermida is
coastal with a longer growing season and
milder winters. The other 14 plantations,
referred to as “satellite tests,” arrayed from
southwest Galicia to northeast Navarra
100 Douglas-fir: The Genus Pseudotsuga
(Figure 4.7), were allocated subsets of the IUFRO coastal Douglas-fir, which originates from that part of
and IFG provenances. The availability of planting its natural range
stock and the nature of planting sites determined
the allocations.
Assessment of growth 5 and 10 years after
plant- ing indicated that the best performing
provenances came from areas between lat 44° N
and 50° N at elevations below 700 m. Their
mother stands are located in regions where
climate is dominated by the Pacific Ocean air
mass. These regions include maritime slopes
along the Georgia Strait in south- west British
Columbia, the Olympic Peninsula, and the Coast
Ranges of Washington, Oregon, and northwest
California. Inland, they include the lower slopes
of the Olympic Mountains and the Coast and
Cascade ranges facing the Puget Trough in
western Washington and the Willamette Valley in
Northwest Oregon.
The height measurements at plantation ages 5
and 10 demonstrated a definite effect of planting
site on growth. It is well illustrated by the two
ma- jor plantations of the 1976 common-garden
test. At age 7 years from seed, mean plantation
height at La Hermida was 69 cm and 39 cm at
Carballa Blanca (Toval 1987). At age 12 from
seed mean plantation height had increased to 233
cm at La Hermida and to 184 cm at Carballa
Blanca (Vega 1990).
Seed source x planting site interaction was
signifi- cant in several plantations. Pairings of test
sites that have many provenances in common
demonstrated them most clearly. Many of the best
provenances showed consistently good growth on
diverse plant- ing sites some of the other best
provenances grew well on one site and poorly on
another (Hernandez et al. 1993). Unfortunately,
future comparisons between the two principal
plantations in the 1976 common- garden test will
be impossible because the test site at La Hermida
was destroyed by fire shortly after the
measurements of growth 10 years after planting
(Vega 1990).
The 10-year results have yielded valuable
infor- mation where seed sources likely to provide
progeny well suitable for planting in northwest
and north- central Spain may be found (Vega
1990). The rapid, juvenile growth of nearly one-
third the number of progenies tested indicated that
where it shows optimal growth, can west Washington (Castle Rock) performed best
successfully adapt even to some of the rather and those from San Juan Island and the Olympic
harsh sites in the mountains of northern Spain.
As Fernandez et al. (1993) stated, “The Pacific
and interior valley climates of western
Washington, western Oregon, and north- ern
California match those of the Iberian
Peninsula so closely that Douglas-fir may
have greater potential in Spain and Portugal
than in the rest of western Europe.”

Italian provenance trials


The experimental plantations established
throughout Italy in the 1920s and 1930s by
Pavari (1958) dem- onstrated that Douglas-fir
could be grown success- fully in the northern
half of the country. The desire to identify seed
sources best suited to Italian needs led to the
initiation of a provenance trial by Pavari at
Vallombrosa Forest, 30 km east of Florence,
in 1951, and was followed by trials at
Acquerino Pistoria Forest in 1954, at
Vallombrosa Forest in 1957, and in Calabria
in 1965 (Morandini 1968).
Pavari chose 10 provenances from
Washington between lat 48°30’ and 45° N
west of the crest of the Cascade Range and
one from northwest Oregon (Vernonia), all
purchased from the Manning Seed Company,
for the 1951 trial. He had originally planned
to compare provenances from north and south
of the 46th parallel, but that was not fea- sible
because seed dealers had made their 1949 and
1950 collections mainly in British Columbia
and Washington. Although intended, progeny
from Italian Douglas-fir stands could not be
included in the trial because cone crops in the
preceding years had been complete failures.
Seed was sown in spring 1951 in the
Vallombrosa nursery, seedlings were
transplanted in spring 1953 and planted as
2−1s in the last week of October 1953 at 700
m in the Vallombrosa Forest (lat. 43°40›N).
Survival was very high for all provenances; it
ranged from 97.5 to 99.5% at age 16 from
seed. The 1967 measurements of height, dbh,
and projections of basal area and volume to a
per hectare basis showed less uniformity in
growth than survival between provenances.
Those from the east slope of the Coast Range
in northwest Oregon (Vernonia) and south-
Chapter 4. Provenance Trials 101
var. glauca, prov- enances formed 3 distinct groups
Peninsula were the least vigorous. The in growth per- formance. Those from Arizona and
provenances from the Washington Cascades were New Mexico
intermediate, except for the provenance Palmer
which belonged to the top performers.
Morandini (1968) noted that the excellent
growth of the provenances Castle Rock and
Vernonia con- firmed Pavari›s opinion that seed
sources in the Coast Range of southwest
Washington and north- west Oregon are most
likely to provide progeny that will adapt
particularly well to the environment of the
Apennines.
Another trial was begun in 1969 by the Istituto
Sperimentale per la Selvicoltura in Arezzo,
Tuscany, with 73 provenances from the 1966/67
IUFRO seed collection (Ducci and Tocci 1987).
They comprised 54 provenances of var. menziesii
from British Columbia, Washington, Oregon, and
California and 20 of var. glauca from Utah,
Colorado, Arizona, New Mexico, and Mexico. In
addition, progeny from 9 Italian Douglas-fir
stands were included in the experiment. Seeds
were sown in late spring of 1969 in the
Vallombrosa nursery and seedlings transplanted
in spring 1971. One group of 2-2 seedlings that
contained all provenances was planted in 1973 in
the Vallombrosa Forest. A second plantation was
established in 1974 at Faltona near Arezzo with 2-
3 seedlings that included half the number of
prov-
enances at the Vallombrosa site.
Measurements taken at ages 5, 7, 11, and 16
from seed in the Vallombrosa plantation
indicated only minor changes in rank of the best
and worst perform- ing provenances of var.
menziesii and the northern subgroup of var.
glauca over the 16-year period. Provenances
with the greatest height and largest dbh at age
16 came from southwest Washington (Castle
Rock) and the western slope of the Oregon (Hebo,
Coquille) and California (Gasquet, Willits, Lower
Lake) Coast Range. Those with the least height
and diameter growth originated from inte- rior
British Columbia (Merritt), the east slope of the
Washington Cascades (Cle Elum) and the east
slope of the California Coast Range (Big Bar,
Weaverville). Growth of the provenances from
the southern subgroup of var. glauca was inferior
to that of var. menziesii. Within the subgroup of
Growth of surviving coastal Douglas-fir was
were the best, followed by those from much better than that of interior Douglas-fir
Colorado; Utah provenances were the poorest. (Ferraris 1993).
The excep- tion from this south to north clinal By 1992, only the plantation at lat 45°05 N at
pattern was the Mexican provenance Saltillo, 330 m elevation in the hills near Turin was left. The
which showed the poorest growth of all
provenances.
Provenances in the Faltona plantation
underwent notable changes in rank during the
first 7 to 9 years but that changed afterwards.
Provenances from the west slope of the
Cascade Range, and the Oregon and
California Coast ranges had consistently the
best height and diameter growth whereas
those from British Columbia, the east slope of
the Cascade Range and the Mt. Shasta region
were always the poorest performers.
Based on the results from the provenance
tests in the 2 locations, considered to
represent the environ- mental conditions in
the north-central Apennines, Ducci and Tocci
(1987) concluded that seed sources between
the coast and the crest of the Coast range in
southwest Oregon, as well as in the northern
California coast Range, appear to be
particularly well suited for the north-central
Apennines. Moreover, the satisfactory growth
of provenances from the Washington and
northwest Oregon Coast range and the west
slope of the Washington and Oregon Cascade
range indicate a broad area of seed origin
optimal for introducing Douglas-fir to the
north- central Apennines.
Another provenance experiment with seed
from the 1966/67 IUFRO collection was
initiated in 1969 by the National Institute for
Woody Plants in Turin (de Vecchi 1973). The
Institute established 3 planta- tions in the
Piedmont in 1970 with 2-0 seedlings from 24
seed sources. They consisted of 15 var.
menziesii provenances, 12 from Washington
and 3 from British Columbia, and also 7 var.
glauca provenances, 3 from Washington and 4
from British Columbia, and progeny from 2
Italian Douglas-fir stands.
Assessments made at ages 2, 6, and 12
from seed showed heavy losses from winter
frost. They ranged for var. menziesii
provenances from 33 to 63%, and for var.
glauca provenances from 17% to 30%. In
growth, the percentages were reversed.
102 Douglas-fir: The Genus Pseudotsuga
plantation at Brosso had been destroyed by fire, and progeny from 14 Douglas-fir stands in eastern
and the one at Voltaggio had to be abandoned
because too few trees had survived. Even in the
Turin plan- tation, survival had decreased to less
than 50% for all provenances, and to less than
20% for 8 of the 15 var. menziesii provenances.
The exceptions were the 2 Italian provenances and
one var. glauca provenance, Revelstoke, from
British Columbia, with survivals of 93%, 60%,
and 67%, respectively.
The Turin plantation had apparently not been
thinned because Ferraris (1993) attributed the ex-
tremely heavy losses between ages 12 and 22 to
competition, winter frost, and physiological
drought caused by a combination of frozen soil
and expo- sure to solar radiation. De Vecchi
(1978) had stated that the generally poor
performance of the North American provenances
was probably a consequence of the use of 2-0
seedlings instead of sturdier trans- plants. The
good results with progeny from the Italian
Douglas-fir stands which also were planted as 2-0
seedlings, suggests poor adaptability to site
conditions in the northwest of Italy as a more
likely reason for the poor performance of the
IUFRO prov- enances in the Piedmont trial.
Austrian provenance trials
Participation in the international IUFRO
provenance trial by the Federal Institute of Forest
Research at Vienna marked the beginning of
provenance tests in Austria. The first two test
plantations were estab- lished in 1973 with 2+2
seedlings. They included 7 coastal and 8 inland
provenances. Seven more planta- tions were
installed with 44 IUFRO provenances in 1977
(Günzl 1981). The number of test plantations had
increased to 53 by 1987 (Günzl 1987).
The observations from these tests (Günzl 1986)
demonstrated that provenances from the west
slope of the Washington Cascades, especially
those that originated from elevations above 500
m, and the southern part of the Olympic
Mountains were the best. Progeny from Austrian
Douglas-fir stands in- cluded in some of the tests
had also shown excellent growth. Inland
provenances, however, grew slowly and suffered
much from late frost.
Schultze and Raschka (2002) analyzed the per-
formance of 177 North American provenances
Austria. The North American provenances seed, provenances from the western Cascades in
stem from IUFRO collections, collections Washington (seed zones 402,403,411) and from
made by Austrian and German foresters, and the coastal regions of Washington and Oregon
seed dealers. Seedlings were raised in the (seed zones 012,053)
Mariabrunn nursery of the Austrian Federal
Forest Research Institute (FBVA) and planted
during 1973 and 1993. Thirteen plantations
were 20 to 25 years old, twelve were 15 years
old, and five were 10 years old when
measured.
Provenances with the best growth came
from the western slopes of the Cascade Range
in Washington and northern Oregon, the
eastern part of seed zone 041 in the Coast
Range in southern Washington, and of seed
zone 052 in the Coast Range in northern
Oregon. Provenances from British Columbia,
most of the Coast Range in Washington and
Oregon, and from the east slope of the
Cascade Range in the two states performed
poorly. The outstanding perfor- mance of
provenances from the southern part of seed
zone 652, the eastern edge of seed zone 652,
and the northwest corner of seed zone 661,
was an exception. The progeny of Douglas-fir
stands in eastern Austria did as well as the
best North American provenances or even
better, except for progeny from mother trees
of apparently interior origin.
Based on the trials’ results, Schultze and Raschka
recommended provenances from the SE seed
zones (Figure 4.8) as best suited for eastern
Austria.

Bulgarian provenance trials


A provenance trial was initiated in 1987
(Petkova 2004). It included 31 provenances of
the variety menziesii, 22 of the northern and 2
of the south- ern subgroup of variety glauca.
The provenances of coastal Douglas-fir came
from the western slopes of the Washington
and Oregon Cascades and coastal regions of
the two states. The provenances which belong
to the northern subgroup of interior Douglas-
fir came from British Columbia, Washington,
Oregon, and Montana. Those which belong to
the southern subgroup of interior Douglas-fir
stemmed from Arizona and New Mexico.
Provenance plantations were established in
five locations in the western half of Bulgaria
(Figure 4.9) in 1989 and 1990. At age 11 from
Chapter 4. Provenance Trials 103

had the best and provenances of inte- 201


rior Douglas-fir the poorest growth. 1010
1020
401 402
600

These early results of the provenance 48°

trial point to western Washington and Mt Vernon


western Oregon as the best seed 011 202 403 621
sources for Bulgaria as has also been Port Angeles
221
shown by existing Douglas-fir stands 212
012
in that country. 222
411
622

Bosnia and 231


Seattle Lake Wenatchee

Herzegovina 412 631


provenance trials Hoquiam
Tacoma

632
Olympia 232
Professor Konrad Pintarić of the 030 421
241
faculty of Forestry, University of Centralia
422 641

C
Naches
242
Sarajevo, ini- tiated an experiment in
46°
1963 with five provenances from

Oc
Washington (Joyce, Wishkah, Elma, 041 430 642
Darrington, Palmer). The provenance AstoriaKelso
440652
651

Palmer is apparently mislabeled as 052


c
653 Goldendale
from Multnomah County, Oregon, 051 042 Washington
elevation 900 m. Pintarić re- ported in Forest Grove
Vancouver The Dalles

Portland
1967 on their nursery per- formance. 661 Oregon
Tillamo-ok

251 451
He outplanted 1-2 seedlings from
662 Maupin
these provenances in 1966 at 053 261 452
Batalovo, about 20 km west of Salem

Sarajevo. The provenance Elma, 671


Warm
Newport Springs672
however, had been replaced by one Corvallis
461
462 463
46°
from Kamloops, British Columbia. 061 263
Waldport 673
Measurements in 1997 (Ballian et al. 473 675 674
471
2003) at age 34 from seed showed the Figure 4.8 Seed zones with provenances recommended for Austria (from Schultze
and Raschka 2002).
standing volume of the four
Washington provenances, projected to
per hectare, ranged from 125 to 235
m3. By contrast, the Kamloops
provenance
had produced much less volume at Was gar, 1091 Yale), and two from Oregon (1099 Pine Grove,
that hing 1100 Grand Ronde). Results were reported
age, namely 135 m3/ha. ton
Pintarić continued trials with some (106
var. menziesii provenances from the 0
IUFRO seed collection with the estab- Sequ
lishment of several experimental im,
plan- tations in 1972. One of the 1069
plantations is located at Crna Lokva Nort
(44° 51′ N, 16°51′ E) elevation 665 h
m. That planta- tion contains two Ben
provenances from Vancouver Island, d,
British Columbia (91029 Thasis, 1090
1036 Alberti), four from Cou
24°
est plantations (squares) and cities (from Petcova 2004).
26°

44°
28°

ROMANIA
ROMANIA

Petrohan
SERBIA
Lesidren
Stara Planina

42° Sofia Kazanluk

l
KjusterndilKostenec
Rilakloster BULGARIA

TURKEY
MACEDONIA

Zlatograd
GREECE 0 60 100 km

060100 mi

F
i
g
u
r
e

4
.
9

L
o
c
a
t
i
o
n

o
f

B
u
l
g
a
r
i
a
n

p
r
o
v
e
n
a
n
c
e

t
104 Douglas-fir: The Genus Pseudotsuga
after 17 (Pintarić 1989) and 32 growing seasons best growth and the fewest losses to winter injuries.
(Govedar et al. 2003). Statistical analysis did The southern
not show significant differences in survival and
volume production between the eight provenances.
Govedar et al. (2003) showed how well these
provenances had performed through a
comparison of their data with those in the
Schober (1987) yield tables for Douglas-fir.
That comparison demonstrated that the volume
production of the provenances in the Crna
Lokva plantation exceeded that given for site I
in the Schober table.
A second plantation is at Gostovic (44°23′ N
18°08′
E) elevation 411 m. Only six of the eight
provenances planted at Crna Lokva were planted
at Gostrovic. The two provenances not represented
are 1069 North Bend and 1091 Yale. Results after
32 growing sea- sons were similar to those
reported for survival and volume growth at Crna
Lokva.
A third plantation at Blinje (43°50′ N, 18°03′
E), elevation 951 m contains only the two
Vancouver Island (1029, 1036) and the two
Oregon provenances (1099, 1100). At this
location too, survival and vol- ume growth did not
differ significantly between provenances at age 32
after outplanting (Ballian et al. 2002).
Although the provenances tested contributed a
very limited sample of coastal Douglas-fir, their
per- formance suggests that progeny from seed
sources in northern Oregon, Washington, and
Vancouver Island is well suited for introduction to
Bosnia and Herzegovina. The large intro-
provenance variance observed in all three tests
provides an opportunity for future selection of
materials for the establish- ment of clone or seed
orchards (Ballian et al. 2003).
Czech provenance trials
The first provenance trials with Douglas-fir in the
Czech Republic were initiated between 1959 and
1961 in three forest districts with seed obtained
from the Manning Seed Company in Washington
(Hofman et al. 1964). Of the provenances tested,
4 came from west-central and southern Oregon, 5
from Washington, 1 from Vancouver Island, and 1
from interior British Columbia (Shuswap Lake).
At plantation age 20, provenances from the
western slopes of the Cascade Range showed the
Oregon and Vancouver Island provenances interior British Columbia had attained
were the poorest performers (Sika 1982). significantly greater heights at plantation age 10
A second set of provenance trials was than those from coastal regions. The latter had
established in 1961 and 1962, one in Bohemia shown a pronounced reduction in annual height
and one in Moravia. These trials included 31 increment after the severe 1972/73 and l975/76
provenances, 5 of which came from Czech winters. The reduction may have been too large
Douglas-fir stands. In these trials a for compensation by subsequent growth.
provenance from Salmon Arm performed Differences in height growth between some of the
best, even better than the best Czech provenances in the 5 plantations also indicated
provenance (Zavadil and Sika 1978). significant seed source x planting site interaction.
A third provenance experiment was The ranking of provenances according to height
initiated in 1968 with 25 provenances from growth had changed little by plantation age 11 in
the 1966 IUFRO seed collection (Sika 1981). the IUFRO experiment and plantation age 20 in
They represented 11 British Columbia, 10 the two earlier sets of provenance trials (Sika
Washington, and 4 Oregon seed sources. Trees 1982).
were outplanted in 1971 and 1972 in five Sika (1981) concluded from the 10-year results of
Bohemian forest districts. Winter drying the IUFRO provenances experiment that 3 regions
(physi- ological drought) and frost damage, should be considered suitable as seed sources for
mainly in the winters 1972/73 and 1975/76 the Czech Republic. In order of preference, they
caused the largest losses. Coastal provenances are the western slopes of the Cascade Range in
from Oregon and Washington suffered most, northern Washington, the lower Frazer River
and provenances from interior British Valley, and the southern inland of British
Columbia and upper elevations in the Columbia. But he cau- tioned that high
Washington Cascade Range were afflicted susceptibility of the interior variety of Douglas-fir
least from these types of climatic injury. to Rhabdocline and Phaeocryptopus needle cast
Provenances from the west slope of the may require reconsideration of the suit- ability of
Cascade Range in northern Washington and inland British/Columbia seed sources.
Chapter 4. Provenance Trials 105

Slovac provenance most promising seed sources for use in Hungarian


trials
The Forest Research Institute at Zvolen initiated Cascades (Seed Zones 402, 412)
the first Douglas-fir provenance experiment in • the rainshadow of the Washington Coast
Slovakia with seed from one of the IUFRO Range (Seed Zone 221)
collections (Tavoda 1991). Test plantations were
• the Frazer River Valley (Seed Zone 1050)
established in 1972 at two locations. The
Kmetova plantation received 21 var. menziesii • the southern inland British Columbia (Seed
provenances from British Columbia, Washington, Zone 2040)
and Oregon, as well as progeny from two Slovak Hungarian provenance trials
Douglas-fir stands. Of the 13 provenances planted
In 1969, the Hungarian Forest Research Institute
at the Velka Straz plantation, 11 belonged to var.
initiated a provenance experiment with 44 IUFRO
glauca and only 2 to var. menziesii. The var.
provenances, 3 from British Columbia, 25 from
glauca provenances originated from Montana (1),
Washington, and 16 from Oregon (Harkai 1983). All
Colorado (6), and British Columbia (4); the 2 var.
Washington and Oregon provenances belonged to the
menziesii provenances came from high elevations
var. menziesii and those from British Columbia to var.
in the Washington and Oregon Cascade Range.
glauca. A plantation with seedlings of the 44
Based on the performance over 20 years in the
provenances was established in 1971 at Zalaerdöd.
field by these provenances, Tavoda (1991)
Both the provenances from Washington and Oregon
provision- ally recommended seed from the
had given very satisfactory results by age 14 from
following regions for use in Slovakia:
seed. Mean height was 6.8 m and dbh 10.7 cm. The
• the west slopes of the Washington
top performers from Washington were the silviculture. The slow-growing provenances from
provenances PeEll, Tenino, Ashford, and Randle. interior British Columbia are considered suitable
Those from Oregon were Vernonia and Estacada. only for the culture of ornamental and Christmas
These results suggested that the west slope of the trees (Harkai 1983).
Washington and northern Oregon Cascade Range
as well as the southern Coast Range of
Polish provenance trials
Washington and the northern Coast Range of Polish provenance trials of northwest American
Oregon contain the co- nifers began in 1960 with an experiment
initiated by Prof. Ilmurzynski of the Polish
Research Institute at Warsaw, which included 5
Douglas-fir provenances from Washington, Idaho,
New Mexico, and prog- eny from 2 Polish
Douglas-fir stands (Bialobok and Mejnartowicz
1970). The institute commenced a second
provenance trial in 1968 with 38 provenances from
the 1966/67 IUFRO collection, 9 commercial
Washington provenances bought from Silva Seed,
and progeny from 9 Polish Douglas-fir stands.
The 38 IUFRO provenances included 12 from
British Columbia, 23 from Washington, and 3
from Oregon. Between 1971 and 1974 eleven
plantations were es- tablished, but only one
contained all 56 provenances (Burzynski and
Gutowski 1973).
An assessment of the performance of the 56 prov-
enances at age 20 from seed led Burzynski et al.
(1990) to the conclusion that all of the
provenances tested can be grown in the parts of
Poland with a moder- ate climate except the
provenance Brookings from southwest Oregon.
Significant differences in growth and frost
resistance between provenances became apparent,
however. Although the interior British Columbia
and eastside Washington provenances showed
inferior growth, they did not suffer any frost
damage. But only 14 of the 28 var. menziesii
provenances from Washington were able to with-
stand the severe environmental conditions without
appreciable cold injuries in the parts of the
country with the most pronounced continental
climate.
An interesting account of the relative
performance of some IUFRO provenances was
provided by Birot and Burzynski (1981). They
compared the perfor- mance of the same 14
provenances, 5 from British Columbia, 6 from
Washington, and 3 from Oregon, at a Polish and a
French site 9 years after planting. The Polish
plantation is at Dolice, lat. 53°14’N, 40 km
southeast of Szczecin. The climate is continental
with oceanic influences because of the vicinity to
the
106 Douglas-fir: The Genus Pseudotsuga
Baltic Sea. Average annual precipitation is 670 reliable estimate of allelic frequen-
mm. The French plantation is at Peyrat-le-
Chateau, lat. 45°49’N in the Limousin. The
location is under the influence of an oceanic
climate and has an average annual rainfall of
1,270 mm.
That comparison showed surprisingly similar
per- formances of provenances at the French and
Polish test plantation although growth was lower
in the Polish test, which reflects the more severe
climatic conditions at Dolice. Provenances from
the foothills of the westside of the Washington
Cascades were the most vigorous, and those from
interior British Columbia the least vigorous. The
only exception was the provenance Brookings
from southwestern Oregon, which grew well at
Peyrat-le-Chateau but poorly at Dolice.
The Institute of Dendrology of the Polish
Academy of Sciences initiated a third trial in 1968
with seed from the 1966/67 IUFRO collection at
Kornik in west- ern Poland. That experiment
initially included 104 provenances from British
Columbia, Washington, and Oregon. Because
excessive frost damage to 4 provenances left
insufficient numbers of seedlings, only 100
provenances are represented in the test plantation
established at Kornik, lat. 52°15’N, in 1971 with
3-year-old seedlings.(Mejnartowicz 1973).
Performance at age 7 from seed already showed
a distinct differences between provenances. The
var. glauca provenances from British Columbia
and Washington showed poor height and diameter
in- crements but excellent cold hardiness. Var.
menziesii provenances from coastal British
Columbia, and especially Vancouver Island,
varied remarkably in growth. But the advantage of
good growth was offset by high susceptibility to
cold injury. Mejnartowicz (1976) concluded on
the basis of the 7-year results that the western
slope of the northern Washington Cascades
appears to contain the best seed sources for use in
Poland.
To determine genetic variation and diversity of
provenances in the Kornik plantation,
Mejnartowicz and Lewandowski (1994) estimated
the allozyme polymorphisms in wind-pollinated
seeds collected from every of the 71 cone-bearing
trees in the planta- tion. He considered one
sample size to be sufficiently large to obtain a
cies at the population scale in a stand. The high cold hardiness. The Oregon and California
trees were 23 years old and represented 41 provenances grew well but suffered considerable
provenances. Expected and observed winter injury in the unusually severe
heterozygosity, proportion of polymorphic
loci, average and effective number of alleles
per locus—indicators of genetic variation and
diversity—were slightly higher than found by
other investigators (Yeh and O’Malley 1980,
Merkle and Adams 1987, Moran and Adams
1989, Li and Adams 1989) in natural stands of
Douglas-fir. Mejnartowicz attributed the
difference to the fact that the test plantation
contains a compressed gene pool which
comprises populations from a vast area within
the natural range of the species. In such an
artificial stand, the probability of mating
among relatives is low. The high level of
genetic polymorphism seems to indicate that
genetic variety was not diminished in the
artificial population.

Danish provenance trials


The first Danish provenance test of Douglas-
fir owes its initiation to the Danish forest
inspector S.M. Storm (1887-1918). He
proposed to establish provenance trials with
conifers from the American West after his
return in 1914 from travels in North America.
Shortly thereafter, with A. Oppermann, he
asked Henry Graves of the U.S. Department
of Agriculture for samples of seed. As a result
of their request, they received, in 1915 and
1916, several lots of seed, to- gether with
information about the parent trees and their
habitats. The shipment included 5
provenances from western Washington, 3
from western Oregon, 5 from California, 1
from eastern Washington, 2 from Idaho, and 2
from Montana.
Oppermann established, in 1918, test
planta- tions with these provenances in mid-
Jutland, Funen, south Seeland, and the isle of
Bornholm. In addition, the plantations
contained a Washington, Oregon, and
California provenance obtained from the seed
firm Johannes Rafn, as well as progeny from
two Danish Douglas-fir stands. Oppermann
reported on the 10-year performance of trees
in the experi- ment. The provenances from
western Washington turned out to be the best
because of excellent growth and relatively
Chapter 4. Provenance Trials 107

winter of 1923/24, in which temperatures dropped


to a low of -27°C. The inland provenances were plantations are given by Lundberg (1957) and
the hardiest but had the lowest rate of growth. Larsen and Kromann (1983). Of the 31
Based on these results, Oppermann (1929) provenances in the 1930 trial, 11 are the same as
advocated the cultivation of Douglas-fir from those in Series II of the Wiedemann trials in
western Washington seed sources but Germany (Table 4.2).
recommended against the use of seed of interior Most provenances tested in the five trials are
Douglas-fir as well as seed from Oregon and from the coastal regions of British Columbia and
California sources. Performance at age 20 of the Washington. Inland provenances were not
1918 test plantations gave further credence to his included anymore after the 1930 trial, except for
opinion. The coastal provenances had attained one in the 1958 trial and 1968 IUFRO trial
heights that ranged from 9.75 to 11.9 m compared because the early tests had already demonstrated
to that slow growth and vulnerability to Rhabdocline
7.0 to 7.9 m for the inland provenances. needle cast made inland Douglas-fir unsuitable for
Moreover, the inland provenances had become cultivation in Denmark. Based on their analysis of
heavily infected with Rhabdocline pseudotsugae the Danish provenance trials, Larsen and
while the coastal prov- enances had remained Kromann (1983) concluded that provenances
healthy (Bornebusch 1939). The 1918 Oppermann from the coastal regions of British Columbia and
trial was followed by five successive trials Washington hold the greatest prom- ise for
initiated by the research branch of the Danish successful cultivation. Within this part of the
Forest Service between 1930 and 1968. They Pacific Northwest, provenances from the south
represent tests of 76 provenances in 17 loca- tions and southwest of the Olympic Mountains were the
throughout Denmark (Table 4.5). Providers of best performers. They considered provenances
seed included the Longbell Lumber Company, from the western slope of the Cascades and the
the Manning Seed Company, and IUFRO. northern part of the Olympic Peninsula suitable
Detailed descriptions of seed sources and location too, but pointed out that they had shown greater
of the test variability in perfor-

Table 4.5 Provenances in five Danish trials from 1930 to 1968 (from Larsen and Kromann 1983) .

Provenances Germination Year of plantation Establishment Number of plantations Seed provider


Colorado 1 1930 1933/1935 4 Longbell Lumber Co.
Montana 2
Idaho 2
Inland British Columbia 5
Coastal British Columbia 3
Washington 12
Oregon 2
California 1
Denmark 1
Germany 2
Washington 5 1937 1940/41 3 USDA Forest Service
Oregon 2 Joh. Rafn
Coastal British Columbia 1 1956 1959 3 Manning Seed Co.
Vancouver Island 3
Washington 3
Denmark 2
Inland British Columbia 1 2 Manning Seed Co.
Vancouver Island 4 1958 1961
Inland British Columbia 1 1968 1971 IUFRO
Coastal British Columbia
11
Washington 3
Denmark 1
108 Douglas-fir: The Genus Pseudotsuga
mance than those from south and southwest of Seeds were sown in spring 1969 in a forest tree
the Olympic Mountains. Larsen and Kromann nursery at the end of the Ulvik fjord, lat. 60°35’N,
recom- mended against the, use of provenances in Hardanger. A test plantation was established
from the southernmost part of Vancouver Island with 2-0 seedlings in the Mobergslien Research
because of high sensitivity to frost. As had area in spring of 1971. The research area is 26 km
already been noticed by Lundberg (1957), growth south of Bergen at lat. 60°10N, long. 5°27›E, at
of progeny from Danish and German Douglas-fir about 100 m elevation. Annual precipitation in the
stands included in the Danish trials was found area ranges from 1,800 - 2,000 mm.
to be on a level with the best American An assessment of survival at age 9 from seed
provenances (Larsen and Kromann 1983). indicated that only 38% of the trees belonging to
coastal provenances were still alive, in contrast to
Norwegian provenance trials
53% of the inland provenances. Most of the losses
Some of the earliest provenance experiments had occurred during the extremely severe winter
with Douglas-fir in Europe were carried out in 1971/72 (Magnesen 1978). In 1981, at age 18
Norway. Børre Giertsen began trials in 1901 in from seed, 34% of the var. menziesii and 48% of
the Ekheng nursery with Douglas-fir from seed the var. glauca trees had survived (Magnesen
sources in Colorado but that work stopped with 1987). Survival of var. glauca provenances from
his death in 1905 (Hagem 1931). Resumption of interior British Columbia above lat. 50°N was
provenance experiments was advocated by higher than that of all other provenances.
Hagem in 1915. He sent Anton Smitt in 1916 to Although these inland provenances had a small
British Columbia to make a seed collection for the proportion of trees with sinuous stems and broad
Norwegian Forest Research Institute. Hagem crowns and had been growing much slower than
tested 24 var. menziesii provenances from British the coastal provenances, Magnesen (1987)
Columbia and Washington, and 6 var. glauca concluded that they “may be the best choices for
provenances from British Columbia. Based on their Norway after all.”
performance in the field for 10 years, Hagem
(1931) concluded that only the northernmost of Swedish provenance trials
the var. menziesii provenances appear to be An initial provenance trial with seed from the
suitable for cultivation in western Norway. 1966/70 IUFRO seed collection was apparently a
Frost killed all trees from Washington seed failure. Most of the provenances originated from
sources. The more frosthardy var. glauca was latitudes too far south to be able to adapt to
deemed to be disadvantageous for western growing conditions in Sweden. Consequently,
Norway because it grows slower than the na- tive seedlings died soon after planting (Martinsson
spruce. Hagem, however, considered Douglas- fir 1990). As a follow-up, another provenance
from the Frazer River area suitable for growth experiment was initiated to investigate the
in eastern Norway. adaptability of provenances from the northern part
Half a century after the initiation of the prov- of the species’ natural distribution.
enance trial by Hagem, the Norwegian Forest In May 1984, Owe Martinsson of the Swedish
Research Institute in Bergen obtained 51 seed lots Agricultural University at Umeå selected 13
from the 1966 and 1968 IUFRO collections. The stands in British Columbia and 2 in northern
ob- jective of the trial with these 51 provenances Washington (Figure 4.10) for cone collections.
was to identify areas that contain seed sources The cone collec- tions and seed extractions were
suitable for West Norway (Magnesen 1973). The made in 1985 and 1986 by two local contractors.
provenances chosen for the experiment were The collections in the Washington and 9 British
thought to have the potential for growth under the Columbia stands were single-tree collections from
climatic condi- tions of West Norway. They 12 dominant trees. In 4 of the British Columbia
included 17 var. men- ziesii provenances from stands, cone collections were made from squirrel
British Columbia, 5 from Washington, and 8 from caches. In addition, seed from bulk collections in
Oregon besides 14 var. glauca provenances from 1 Danish and 2 Swedish Douglas-fir stands were
British Columbia and 7 from Washington. included in the experiment.
Chapter 4. Provenance Trials 109

68°

66°

0100 km

64°

Trekilen

Storsela Bergdal
62°

Vikarbyn

60°

56°

58°
136°

Bruzaholm
Tönnersjöheden
Mästerby

2. Takla Lake 56°

52°
I Fort James
Prince George

12. Bella Coola 4. McBride


132°
Quesnel
Horsefly
IV
10. Horn Lake
II 8. Blue River
13. III
128° Campbell River 5. Sicamous
9. Shushwap

0100200 km

14. Hoh Valley 15. Darrington

124°
120°
116°

Figure 4.10 Collection sites for Swedish provenance tests (open circles) and sites of test plantations (closed
circles); from Martinsson and Kollenmark (1993).

Seedlings were raised in the Almfors tree 1 year after planting on the sites in central Sweden
nursery in Hälsingland and planted as 2-year-old (Martinsson and Kollenmark 1993). Survival
seedlings in 3 test plantations in southern Sweden ranged from 71% to 100% in southern Sweden,
in 1990 and in 4 test plantations in central Sweden but coastal provenances had the lowest rates of
in 1991 (Figure 4.10). Survival of seedlings was survival in all three plantations. The survival rate
recorded 2 years after planting in the southern was consider- ably lower in the central Swedish
plantations, and than in the south
110 Douglas-fir: The Genus Pseudotsuga
Swedish plantations. Survival for the 10 best southwestern archipelago. Hagman (1973)
prov- enances, which excluded all coastal emphasized that the
provenances, varied between 49% and 84%.
Although Martinsson and Kollenmark (1993)
stated that 2 years in the field are too short a
period for an evaluation of the seed sources in the
experi- ment, they nevertheless pointed out that
coastal provenances had such a low rate of
survival that their use anywhere in Sweden would
probably be doomed to failure. For introduction to
central Sweden, only provenances from the
northernmost part of the spe- cies’ natural range
or from high elevations in the interior of British
Columbia should be considered. A look at the
map (Figure 4.10) makes the reason for these
recommendations very clear.
Finnish provenance trials
Small test plantations established in Finland
during the first half of the 20th century had shown
generally poor performance by both varieties of
Douglas-fir. Some variety glauca provenances from
interior British Columbia, especially the upper
Frazer River valley (Heikinheimo 1956), were an
exception. Because early trials lacked adequate
replication, the Finnish Forest Research Institute
decided to initiate another provenance trial with
an up-to-date experimental design, when seed from
the IUFRO collection became available. But in
view of the earlier experience, their choice of
provenances was limited to the 9 north- ernmost
(IUFRO numbers 1001–1008) of the IUFRO
collection in British Columbia (Hagman 1973).
Seeds were sown in May 1970 in a tree nursery
on the island of Nagu (lat. 60°11’N) in southwest
Finland. Frost in fall and winter of the first year
in the nursery killed nearly all the seedlings from
the 2 var. menziesii provenances in the
experiment, but the 7 var. glauca provenances
suffered far less dramatic losses. At the end of the
third year in the nursery, survival of seedlings
from the 7 interior provenances ranged from 48%
to 65%.
Although the results of the trial with IUFRO
provenances covered only 3 years in the nursery,
they appeared to give credence to previous experi-
ence that the variety menziesii seems to be unable
to adapt to Finnish conditions, even under such a
mild climate as prevails in the country’s
number of provenances in the IUFRO Several trials were begun during 1930 - 1932
experiment is small but expressed the hope by Prof. Mathiesen in the forest of Tartu (formerly
that it can give some guidelines for further Dorpat) University. Results confirmed the
selection. suitabil- ity of var. glauca seed sources from
interior British Columbia for Estonia (Margus
Latvian provenance trials 1963). In the spring of
The Latvian Research Institute for Forestry
Problems has embarked on a large program of
provenance trials. The trials involve 300
provenances, 125 of which are from the
IUFRO collection. The other 175 provenances
come mostly from second generation
Douglas-fir stands in Russia, the Ukraine,
Bulgaria, Poland, Hungary, Norway,
Denmark, Germany, and the Netherlands
(Pirags 1968, 1990). Initial results indicated
that provenances suitable for Latvia come
from the area between lat 45° and 55° N, and
long 117° and 125° W. Progeny from stands
in the Baltic region (Estonia, Latvia,
Lithuania) did very well. Provenances from
south of lat 43° N are absolutely unsuitable
for Latvia (Pirags 1979, 1990).

Estonian provenance trials


Among the earliest provenance experiments
in Europe is that of Count Berg, established
on his es- tate in Sagnitz, Estonia. After
disappointing results with seed purchased
from European seed dealers, Berg was able
to obtain 13 seed lots in 1909, and another
12 in 1910, of known origin from the US
Bureau of Forestry. All were var. glauca
provenances from New Mexico, Colorado,
Montana, Idaho, and the east slope of the
Washington Cascades, except for a
Snoqualmie, Washington, and Lake Tahoe,
California, provenance. Initial results
suggested that var. glauca provenances
would be more suit- able for use in Estonia
than would var. menziesii provenances
(Berg 1912, Zon 1913). Some of the trees
from this experiment were still alive in
1978, but records pertaining to the
experiment are unfor- tunately lost (Etverk
1978). M. Sievers, chairman of the Baltic
Forestry Association, bought seed of
different provenances directly from North
America in 1911 but records of plantations
established with that seed are also lost
(Margus 1961).
Chapter 4. Provenance Trials 111
between lat 47° and 48° N. Other var. menziesii
1973 a provenance test was begun in the state provenances had begun to perform less satisfactorily.
forest managed by the Estonian Academy of All var. glauca provenances had
Agriculture at Järvselja. Provenances from
northern interior British Columbia and Montana
did not suffer frost damage in the first 5 years
since germination but grew slower than progeny
from Estonian Douglas- fir stands (Etverk 1978).
Initially, Estonia was unable to participate in
the international IUFRO Douglas-fir provenance
experi- ment. In spring 1978, however, H. Barner
provided 23 lots of seed from British Columbia,
Washington, and Oregon from the IUFRO seed
collection. They were sown in the plastic
greenhouses of the Estonian Forest Research
Institute (Etverk 1978). Information about the
progress of this experiment was not avail- able to
us.

Turkish provenance trials


Except for a few small Douglas-fir plantations es-
tablished after World War II, experience with the
species’ adaptive ability to Turkish site conditions
was lacking. To provide a basis for the selection
of suitable seed sources, the Research Institute for
Poplar and Fastgrowing Conifers at Ismit initiated
a comprehensive provenance experiment in 1971,
with 85 var. menziesii and 33 var. glauca
provenances from the IUFRO seed collections
(Simsek 1978).
Seeds were sown in 1972 in the Alendag
nursery near Istanbul. Because 37 of the
provenances had a low percentage of germination,
only 81 provenances were available for
establishing 9 plantations on eco- logically
different sites along the Black Sea coast in 1974.
Latitude of the sites ranged from 40°44’ N to
41°23’ N, longitude from 29°48’ to 38°25 E,
elevation
from 25 m to 1,340 m.
An assessment of performance at age 6 from
seed, based on survival, height, and diameter
growth, showed provenances from elevations
below 600 m of the west slope of the Washington
and Oregon Cascade Range, and the east side of
the California Coast Range, to be the best ones
(Simsek 1978). Subsequent assessments at ages 9
and 11 from seed (Simsek 1980, 1982) revealed
excellent growth by provenances from the
Washington central Cascades and Coast Range
shown particularly poor growth and could be
ruled out for further consideration for use in
Turkey.
After 14 growing seasons in the field,
marked changes in performance between var.
menziesii prov- enances became apparent
(Simsek 1987). Growth of British Columbia,
Oregon, and California provenanc- es had
greatly slowed, and their rate of survival had
decreased to unacceptably low. Simsek (1987)
was confident that the performance of the
provenances from the central Washington
Cascades in both low- and high-elevation test
plantations suggested that the central
Washington Cascades contain the most
promising seed sources for the Black Sea
region. He justified his opinion with these
words, “if adaptabil- ity of provenances is
looked upon as optimization between good
growth and a high rate of survival, than the
provenances from the central Washington
Cascades show the greatest adaptability,
among the provenances tested.”

Taiwan provenance trials


The Taiwan Forest Research Institute initiated
a provenance experiment with P. menziesii to
explore the possibility of successfully
introducing the species to the island’s forests in
the early 1970s (Yang 1978). Seed collections
were made in 14 stands in coast ranges of the
Pacific Northwest: 9 in California, 3 in
Oregon, 1 in Washington, and 1 on Vancouver
Island Seeds were sown in the Chi-tou and
Chu-yun Shan nurseries. Growth of
provenances from south- ern or lower
elevation origins was consistently supe- rior to
that of provenances from northern or higher
elevation origins. Based on their performance
in the two nurseries, Yang considered four
provenances from the fog belt of California to
be the most adapt- able to the environment of
medium high elevations in Taiwan. The
locations of origin of these four prov-
enances are given in Table 4.6.

Table 4.6 Locations of the origin of four provenances used in


Taiwan trials.
Provenance Latitude N Elevation (m)
A 37°13’ 394
B 39°38’ 380
D 40°23’ 454
E 41°78’ 563
112 Douglas-fir: The Genus Pseudotsuga

New Zealand provenance diameter, wood density, heartwood percent, and


trials
Second- and third-generation stands of Douglas- survival) were measured or scored on 15 prove-
fir in New Zealand, in contrast to those in Europe, nances in six plantations (Figure 4.12), and on all
have provided far more seed than imports from 44 provenances in three of the six plantations. The
North America. Location of the original seed 15 provenances were the same ones already
sources of stands established before 1926 is judged at age 5 in the field by Thulin (1967) as
uncertain but a record exists of site of seed origin the most prom- ising. Of the traits analyzed,
used in plantings after that date (Wilcox 1978). height varied the most and gave the clearest
Although seed source descriptions were available separation of provenances. Genotype x
for the post-1926 imports, performance of their environment interaction for all traits was small,
progeny was not evaluated by provenance tests. indicating that the best provenances were the
Preparations for provenance trials commenced superior performers on all sites. The best
with collections from the 1955 cone crop in provenances came from low elevations in the fog
British Columbia, Washington, Oregon, and belt of northern California and southern Oregon.
California made by American seed suppliers Washington provenances and those from the
(Anonymous 1994). Plantations with seedlings Sierra Nevada of California grew distinctly slower
from the North American collections were at all test sites than the provenances from the
established at 10 loca- tions on the North and California and Oregon coast.
South Island. Included in that trial were also As a result of the 1959 trial, seven provenances
progeny from New Zealand Douglas-fir stands at (Table 4.7) were identified as those from whose
Dusky, Kaingaroa, Whaka, and Tapavera. After sources of origin seed should be imported for use
the sixth growing season in the field, Washington
in New Zealand and on which a breeding program
provenances from low altitudes were generally the
should be based (Wilcox 1974). Six of these seven
tallest at all locations. The four New Zealand
provenances had ranked at the very top in height
provenances were among the best at most sites
growth at age 5 in the field. In the words of
(Wilcox 1978).
Wilcox (1974), “This early assessment provided
A second trial was begun with 44 provenances
the informa- tion which led to the importation of
from Washington, Oregon, California, and a bulk
several com- mercial seed lots from coastal
seed lot collected at Kaingaroa Forest. The North
Oregon and California. The 13-year assessment
American collections were made in 1956 by Egon
results confirm that these interim choices were
Larsen (Figure 4.11) to fill out gaps in the first
substantially correct.”
trial (Anonymous 1994). Most of his seed lots
The New Zealand tree seed company
were from cone collections on at least 10 trees.
PROSEED funded, in 1988, the selection of plus
Seedlings for the second trial were planted in
trees from the best coastal fog belt provenances in
1959 at 19 locations in the North and South Island
(Sweet 1964). the 1959 trial. Because the test plantations had
In the assessment of performance 13 years af- been thinned in 1976 only about 20 of the original
ter outplanting nine traits (height, diameter, stem 144 trees per plot were left. All trees in the best
straightness, malformation, needle retention, provenances were measured over six sites (Table
branch 4.7) and the best tree per plot selected. The
selected trees were grafted into a seed

Table 4.7 Superior Douglas-fir provenances selected for use in New Zealand (from Wilcox 1974).
Provenance Origin Characteristics
636 Oregon, Deadwood Excellent form, and good needle retention, vigorous
641 Oregon, Four Mile (Bandon) High wood density, vigorous
642 California, Berteleda (Gasquet) Very vigorous and excellent form
647 California, Mad River (Korbel) Very vigorous and high wood density
654 California, Caspar (Fort Bragg) Vigorous
659 California, Stinson Beach High wood density, vigorous and excellent form, prone to needle cast
660 California, Santa Cruz Very vigorous, low wood density, prone to needle cast
Chapter 4. Provenance Trials 113

50°

British Columbia
36°

AucklandRotorua ConservancyConservancy
578 Rapunui
Kaingaroa
631
579

586
584 581Washington Gwavas
Oc

45° 576
583 632
575
634 41° Nelson Conservancy Wellington Conservancy

636 585
635 637 633 Golden Downs
580 577 Westland Conservancy
641 640 Hanmer
639 536 NEW ZEALAND
Canterbury Conservancy South P acific Ocean
642
Oregon
643
647
644
646 645
40° 603
648 Southland Conservancy
650 653
649 652 Rankleburn
654 46°
655
658 657 90090180270360 K

659656
San Francisco
660
170°
174° 178°

1967 with the establishment of test plantations at


California
35°
Kaingaroa, Rotorua, and Gwavas. That experiment

Los Angeles

160 0 160 320 K

Figure 4.11 Sites of Egon Larsen seed collection (circles);


provenanc- es recommended for use in New Zealand based on
performance 16 years after outplanting (triangles); from Wilcox
(1974).

orchard at Canterbury. In addition, PROSEED and


the New Zealand Forest Research Institute funded
a seed collection from southern coastal California
and coastal Oregon with the primary objective of
widening the genetic base to include previously
untried provenances (Anonymous 1994, Miller
and Knowles 1994).
A third provenance experiment was begun in
Figure 4.12 Location of six plantations of the 1959 provenance test because of infec- tion with Phaeocryptopus
in New Zealand (from Wilcox 1974).
gaeumannii. P. flahaulti, with an average height
included two Mexican Douglas-firs, of 3.7 m, was as vigorous as the 3.5-m-tall
Pseudotsuga flahaulti and P. macrolepis, a Kaingaroa provenance. The second Mexican
California provenance from Santa Cruz, and Douglas-fir, P. macrolepis, had only attained a
New Zealand Douglas-fir from Kaingaroa height of 2.6 m and showed poor vigor.
(Wilcox 1978). The assessment of the 7-year Additional provenance tests were established
performance of the trees at the Rotorua site in 1971 and 1974 on numerous sites throughout
showed that the Santa Cruz provenance, with New Zealand. Their results have led to the
an average height of 4.6 m, was significantly identification of distinct, superior local strains. A
taller than the New Zealand and Mexican 1980 report of the New Zealand Forest Research
Douglas-firs but had suffered loss of needles Institute at Rotorua (1981) describes four such
strains as follows:
114 Douglas-fir: The Genus Pseudotsuga

Fort Bragg strain shelterbelts in the Tapawera district near Golden


The native provenance of this strain originated in
Jackson State Forest close to Caspar, Fort Bragg,
California (altitude 160 m). It is typical of the
low- elevation coastal Californian provenances
that have grown so well in provenance tests,
particularly in the North Island. Seed stands at
Rotoehu Forest and Golden Downs Forest have
been formed in blocks planted in 1959 and so far
88 kg of seed have been collected. Compared to
the widely used and more familiar Kangaroa
strain of Douglas-fir (originally from Figure 4.13 Australian Douglas-fir provenance plantations (from
Washington), the Fort Bragg strain is distin- Griffin and Matheson 1978).
guished in the nursery by its exceptional vigor,
comparatively early flushing in spring, and by its
bright green color.
Ashley strain
This strain has proved to be a consistently good
grower on both low- and higher altitude sites in
the North and South Islands. A seed stand has
been cer- tified at Ashley Forest, and new second-
generation seed stands are being developed at
Golden Downs and Ashley. The Ashley strain
originated from
Downs Forest, Nelson. The native origin of these
early introductions is not known; in flushing time,
however, the Ashley strain (and other seedlots of
Tapawera ancestry) is intermediate, similar to that of
provenances from coastal localities in Oregon.
Beaumont strain
The origin of this strain (Beaumont Forest, Southland)
can be traced to various stands of Douglas-fir in the
Tapanui district, supposedly of Washington origin.
The strain has not yet been performance-tested but is
nonetheless expected to be reliable for planting in the
South Island.
Kaingaroa strain
Most of the Douglas-fir in Kaingaroa Forest origi-
nated in the state of Washington. “Kaingaroa strain”
refers to any of the larger seedlots collected from
various compartments in the forest. It is a depend-
able, late-flushing strain, giving maximum protec-
tion against late frosts, but its growth rate is not
exceptional.
Australian provenance trials
The CSIRO Division of Forest Research initi-
ated in 1970 a provenance experiment with 40
seedlots from the 1966/69 IUFRO collection and
some non-IUFRO seedlots from low-elevation
AUSTRALIA
sites in the California coastal region. The IUFRO
seedlots included Pseudotsuga menziesii var.
menziesii from British Columbia, Washington,
Oregon, California, and Mexico (Griffin and
Matheson 1978). Nine test plantations were
established in 1972-73 in New South Wales,
New South Wales 30°S
Victoria, and Tasmania on sites considered to 9

have potential
n
for the growth of Douglas-fir.
These sites cover a latitudinal range from 31° 7, 8 to
35°S
41°S (Figure
Kamona 4.13) and elevations from 1806to
Victoria
Star of Peace
1,200 m. Because a common design
Smiths Plains
4 5 for all

plantations was not attempted, only 11 of the 40


Narbethong
Mansfield
seedlots
Bondiwere planted at all 9 sites. With the40°S
Buccleugh
exception of one Washington provenance
Buccleugh 3 2
1 (Cle
Nundle Tasmania
Elum), the provenances represented in each of
the test plantations originated from the southern
Oregon and California coast ranges.
Measurements of height growth at plan- tation
age 5 clearly demonstrated the faster
Chapter 4. Provenance Trials 115
and AFOCEL in 1979 (Breidenstein et al. 1990). In
growth of provenances from low elevations in the
coastal regions of southern Oregon and northern
California relative to those from comparatively
short distances farther inland in the coast ranges.
But Griffin and Matheson (1978) cautioned that
these early results, “do not demonstrate clear
superior- ity of any particular seed collection area
within the low-elevation coastal areas of northern
California, Oregon, and Washington, and it may
be that traits other than growth potential per se
would influence the final choice of provenances”.
They concluded, however, that—based on current
evidence—seed from sources in the coastal fog
belt of northern California and southern Oregon
would be recom- mended for use in Australia.

What has been learned


from provenance
experiments?
Provenances experiments have shown that the va-
riety glauca, with perhaps a few exceptions, is not
desirable for introduction outside its natural
range, mainly for two reasons: (1) the growth of
inland Douglas-fir is inferior to that of the variety
menziesii, and (2) inland Douglas-fir is highly
susceptible to infection by Phaeocryptopus
gaeumannii (the Swiss needle cast pathogen),
which significantly reduces growth and frequently
leads to premature death.
Because of the generally poor performance of
variety glauca in areas of introduction, emphasis
has focused on variety menziesii in provenance
studies. “Studies initiated in the first half of the
20th century showed that the most promising seed
sources for many parts of Europe may be found in
western Washington and northwestern Oregon.
The experi- ence gained from these early studies
also pointed to the need for a more accurate way
to assess the performance of a given provenance
in different geo- graphic locations of introduction.
That led to the de- cision by IUFRO Working Party
S2.02-05 Douglas-fir provenances at its 1978
Vancouver, British Columbia, meeting, to create a
database that would permit a valid comparison of
the results of field tests obtained by participants in
the international IUFRO Douglas- fir provenance
experiment
The database was set up in France by INRA
and the crest of the Cascade Range; and
September 1989, the base contained data “interior” that provenances were from east of
from 108 test sites provided by 20 the crest of the Cascade Range. The database
institutions in 14 European countries and had been originally designed to compare
Canada. Those data represented barely half of performance of provenances based on
the 33 countries that had received seed from assessment of three traits: survival, height
the IUFRO collections. Breidenstein et al. growth, and time of budburst. Bud burst,
(1990) evaluated available data by however, had to be omitted from consideration
arranging plantation sites into these because of insufficient data. Provenances
ecological groups: (1) sites with continental belonging to the southern subgroup of variety
climate in north-eastern Europe; (2) sites glauca had to be excluded from analysis
with a mild oceanic climate in northwestern because they were tested in few of the sites.
Europe and southwestern British Columbia;
(3) sites exposed to a relatively harsh Survival
oceanic climate with low mean annual Survival at or close to age 10 in the 108 test sites
temperatures in northwest- ern British ranged from 70% to 85%. Provenances within
Columbia, Norway, and a few locations in groups 2 and 4 had higher rates of survival than
France and western Spain; (4) southern those within groups 1 and 3, however,
Europe south of lat 48° N. Provenances provenances originating from elevations above
were separated into three groups designated 1000 m had slightly higher survival than those
as coastal, intermediate, and interior, based from lower elevations. In group 1 sites, coastal
on broad geographic areas of origin: provenances from southern Oregon had the lowest
“coastal” indicated that provenances origi- survival; in group 4 sites, provenances from
nated from the area between the Pacific interior British Columbia had the lowest sur-
shore and the crest of the Coast Range; vival. Provenances from low elevations in western
“intermediate” that provenances came from Washington had consistently the highest survival
the area between the crest of the Coast Range on sites in groups 1,2, and 4.
116 Douglas-fir: The Genus Pseudotsuga

Height Mediterranean region of France. Most Douglas-fir


growth
Analysis of the influence of latitude of origin have
showed a parabola-like pattern of variation of given an entirely unbiased picture because of differ-
growth with latitude of origin for provenances ences in data collection and experimental procedures.
from the coastal area. Growth of provenances Breidenstein et al. (1990) attributed the surprisingly
from the intermediate area showed a linear high rate of survival to the choice of favorable test
increase with origin from lat 39° to 49° N. Height sites, focus on subsets of provenances deemed well
growth of interior provenances that came adapted to the experimental site and submission of
primarily from the northern part of their range data to the base only from successful experiments.
failed to reveal a relation with latitude. The largest Whatever bias may have been involved in the data
number of fast-growing provenances came from analysis, its results largely confirm the findings of
el- evations below 600 m in Washington and provenance tests established before the international
Oregon. In contrast, provenances from low IUFRO Douglas-fir provenance experiment.
elevations in British Columbia had moderate
height growth. IUFRO collected few provenances
Southern hemisphere
from low-elevation seed sources in California; Participants in the international IUFRO Douglas-fir
they were planted in very few test sites. So, provenance experiment from Australia and New
California is represented mainly by relatively Zealand did not provide results from their tests to the
slow growing provenances; that stem from database set up by INRA together with AFOCEL.
elevations above 900 m. Although direct comparisons between the perfor-
The data submitted to the database may not mance of individual provenances in the northern and
southern hemisphere are therefore not pos- sible, provenance studies have shown considerable
at least one notable difference is apparent. varia- tion not only between but also within
Provenances originating from the coastal regions populations. That may partially explain the broad
of southern Oregon and northern California adaptability of trees of many provenances
performed extremely well in Australia and New (Kleinschmit and Bastien 1992). Individual trees
Zealand, in contrast to Europe, where these can also have rather broad adaptability, but,
provenances were generally failures, except for a considerable differences in individual adaptability
few locations in the may be explained by differ- ences in
heterozygosity. As Li and Adams (1989) have
indicated, expected heterozygosity is highest in
the Pacific coastal regions of northern Oregon,
Washington, and Vancouver Island. Kleinschmit
and Bastien (1992) suggested that might be an ad-
ditional explanation for the broad adaptability of
provenances from this part of the Pacific
Northwest.

Progeny of introduced populations


Provenance experiments have shown that progeny
of some introduced populations of Douglas-fir
per- formed as well, or even better than the best
native North American provenances. As the origin
of most of the populations introduced before the
second half of the 20th century is unknown, their
progeny is sometimes referred to as “artificial
stands” or “land races.” Attempts have been made
to identify their provenance. Berney (1972) made
probably the earliest of such attempts by using the
DNA content of embryo cells to determine the
origin of a Douglas- fir stand in Switzerland. In
studies that tried to trace the North American
parents of Douglas-fir stands in Switzerland
(Stauffer and Adams 1993), France (Prat and
Arnal 1994), and Germany (Klumpp 1999)
isozymes were used as genetic markers. Rehfeldt
and Gallo (2001) tried to determine the parentage
of a Douglas-fir stand in Argentina by using
quantitative traits because of their suitability for
estimating ge- netic variances. Results of all these
studies indicated broad rather than specific
geographic locations from where the progenitors
of the Douglas-fir land races had come from.

Conclusion
Since most of the classical questions of
provenance research have been answered for
Douglas-fir the focus is now on breeding and gene
conservation.
5. Tree Breeding and Improvement
Richard K. Hermann

J
akob Roeser (1926) foresaw the need for mountains at the south- ern part of the region is for best
growth of Douglas-fir
tree improvement almost 90 years ago: “The
im- provement of forests by any mean
whatever is
at present so urgent, and a creative method makes
so great an appeal to the American type of mind,
that it is desirable to direct the attention of forest-
ers generally to the possibilities of tree breeding.”
Genetic improvement in forest trees mostly comes
from increasing vegetative growth, improving re-
sistance to biotic or abiotic stresses, and
enhancing wood and stem quality (Howe et al.
2006). Tree breeding efforts in western North
America have centered on Douglas-fir in
particular because of its great economic and
ecological importance. The species exhibits high
levels of genetic variation for all economic and
adaptive traits studied, providing a rich
foundation for genetic improvement (Howe et al.
2006).
The climate within the range of coastal
Douglas-
fir changes dramatically from west to east. A
narrow coastal strip extending from San Francisco
to north- ern Washington has an unusually long
frost-free period and annual rainfall of up to 4400
mm (US Department of Commerce, Weather
Bureau 1957). The change is due to mostly north-
south mountain ranges where adiabatic cooling
and rain-shadow drought effects interact. More
climatic complexity is added each growing season
by droughty summers in the Douglas-fir region,
particularly below lat 42° N. Silen (1989) noted
how changes in cold and drought with elevation
and latitude affect growth patterns:
Typically rain-shadow drought is most severe in
valley bottoms and decreases with increasing
elevation. This is the opposite in direction to the trend
of decreasing growing season length with elevation
caused by diabatic cooling. A typical situation in the
to occur at middle elevations, with growth like in matching genetics to environment (Silen
restricted by increased cold upward and by
more intensive drought downward. Along the
1989). McArdle and Meyer produced their tables
west slope of the Cascades, the corresponding in 1930, when the Douglas-fir region was mainly
elevation of maximal growth descends in natural forests. The tables give total yield by
northward from mid-elevation in southern
Oregon to about sea level in northern
decade for stands indexed into 3-meter height
Washington, with a gradu- ally narrowing band classes. For example, dominant trees on best,
of droughtiness below and colder temperatures average and poorest sites (site classes I, III, and
upward toward timber line. (Silen 1989)
V) attain 61, 43, and 24 m in height, respectively,
For both varieties, low temperatures are the at age
major limiting factor within the northern 100. They sampled over 1,900 plots throughout the
part of the range, whereas lack of moisture is region, confining plots to pure stands of “Normal”
the predominant limiting factor in the stocking, which were dense-canopied in the self-
southern part. thinning stage and generally almost devoid of
Tree species occupying widely varying other vegetation. Moreover, when analyzed
habitats can be subdivided into geographic decades later in terms of −3/2 power law of self-
races (Callaham 1970). Evidence has shown thinning, the data produced slopes with an
that these races are the re- sult, in large part, exponent value of −1.5 the theoretical limiting
of adaptation to the environments in which value (Silen 1989):
they are found (Adams 1981). The Douglas- A frequently overlooked, but important point is that
fir yield tables of McArdle and Meyer (1930), the 3/2 power law, which expresses average tree size
over the range of spacing, has an interchangeable,
for instance, indicate that feedback alternate
mechanisms are precise, probably template-

117
118 Douglas-fir: The Genus Pseudotsuga
form, the law of constant final yield. That version Patterns of genetic variation in adaptive traits
uses the same data to express volume per unit area
over spacing instead of average tree size over spacing, have been studied using long-term field tests and
has an interchangeable, alternate form, the law of short-term experiments in outdoor nurseries,
constant final yield. That version uses the same data green- houses and growth chambers. These
to express volume per unit area over spacing instead
of average tree size over spacing. If data from fully- studies have confirmed the overriding importance
stocked stands fit the theoretical -1.5 slope, it follows of temperature and moisture regimes in shaping
that a given site will also attain a constant final yield genetic variation throughout the range of
over a great range of initial densities. It is important to
grasp that most fully-stocked stands in the sample Douglas-fir. The species is particularly responsive
must have approached theoretical maximum yields to these selective forces, and is considered an
which occur at the theoretical -1.5 slope. To have adaptive specialist (Howe et al. 2006).
such sampling accurately approximate theoretical
yield maxima suggests that natural popula- tions at St. Claire et al. (2005) described and mapped
each locale must be attuned phenologically to use the pat- terns of genetic variation in adaptive traits in
entire growing season. (Silen 1989) coastal Douglas-fir in western Oregon and
Washington. They measured growth and
Genecology of Douglas-fir phenology in seedlings grown from wind-
pollinated seed of 1338 parents in naturally
The term genecology was coined by the Swedish
regenerated stands at 1048 locations. Seedlings
botanist Turesson in 1923. As cited by Langlet
were measured for traits of emergence, bud
(1979,
phenology, growth, and partitioning. The au- thors
p. 657), Turesson wrote: “It seems appropriate for
concluded that adaptation of Douglas-fir popu-
several reasons to denote the study of species-
lations to Pacific Northwest environments appear
ecolo- gy by the term genecology (from the Greek
to be largely a consequence of trade-offs between
‘genos,’ race, and ‘ecology’) as distinct from the
selection for traits to avoid exposure to cold and
ecology of the individual organism, for which
traits that confer high vigor in mild environments.
study the old term autecology seems to me to be
Winter temperatures and frost dates are of great
the adequate expres- sion.” Turesson additionally
im- portance to population differentiation.
defined genecology as “the study of the species
Selection for drought avoidance by early budburst
and its hereditary habitat types from an ecological
also appears to have resulted in population
point of view” (Langlet 1979, p. 657).
differentiation. The authors stated that an
The genecology of Douglas-fir has been
important question arising from their work
studied at many scales, using wide ranging
remains unanswered: what specific genetic and
provenance tests (Wright et al. 1971) to studies of
epigenetic phenomena are responsible for
variation within a single watershed (Campbell
geographic variation observed in adaptive traits?
1979). Studies of gene- cology focus on “adaptive
To address this fundamental question, parent trees
traits” – traits believed to be under strong natural
from this study are currently being genotyped at
selection because they confer adaption to the
candidate genes presumably involved in cold
environment and enhance individual fitness
hardi- ness and drought tolerance.
(Howe et al. 2006). Studies of adaptive traits
typically include survival, height growth, fall and Quantitative genetics and inheritance
spring frost hardiness, drought hardiness, vegeta- As understanding of genetic and environmental
tive bud phenology (the time of bud set in fall and variation is important for designing breeding strat-
budburst in spring) and the frequency of second egies, picking suitable mating designs, designing
flushing (Howe et al. 2006). Second flushing field tests, and predicting genetic gains. Key
occurs when a tree stops elongating and sets a pieces of information include relative amounts of
bud, then flushes a second time in the same addi- tive vs. non-additive genetic variance,
growing season. Douglas-fir exhibits high levels genetic and environmental variances,
of genetic variation for each of these adaptive heritabilities, and genetic correlations (Namkoong
traits within and among varieties, provenances, and Kang 1990). Most infor- mation on quantitative
and populations (Silen 1978, Rehfeldt 1989). genetic parameters is derived
Chapter 5. Tree Breeding and Improvement 119
diameter have low heritabilities but
from analyses of wind-pollinated families
collected from natural populations as compared to
advanced- generation breeding populations (Howe
et al. 2006). Additive genetic variation, the
variation associ- ated with the additive gene
action effects, is the main reason that progeny
resemble their parents and the main determinant
of population response to selection (Johnson
1988). Douglas-fir breeders are mainly interested
in the additive genetic variance because most
breeding strategies rely on improving populations
via recurrent selection, and because most
materials are produced via wind-pollinated seed
orchards, which do not capture the non-additive
component (Howe et al. 2006).

Heritabilities and amounts of


genetic variation
One of the most important indicators of potential
breeding success is heritability (h2), a measure of
the relative degree to which a character is
influenced by heredity as compared to
environment, that is the ratio of genetic variation
to phenotypic variation. The higher the
heritability, the more an individual’s phenotype is
indicative of its genotype (R Johnson 1998). Trait
heritabilities have received much atten- tion
because they integrate information on genetic and
environmental variation, and because they can be
altered to increase genetic gains, primarily by re-
ducing environmental variability in genetic tests
and by increasing family size to increase family
heritabili- ties (Howe et al. 2006). These authors
compiled from 23 sources a list of mean
heritabilities for common traits measured in
Douglas-fir progeny tests under field conditions.
They included only experiments in which trees
were at least 4 years old.
The list shows that heritabilities for bud break,
bud set, second flush, spring cold hardiness,
branch angle and wood density (specific gravity)
are mod- erate to high; and low to moderate for
growth traits (height and diameter), fall cold
hardiness, stem de- fects (ramicorn branches,
forks and sinuosity). Traits vary, however, in their
degree of genetic control (h 2) and the relative
amount of genetic variation. Traits such as wood
density have high heritabilities but low genetic
variation whereas other traits such as height and
genetic variation (Howe et al. 2006).
high levels of variation (Cornelius 1994). Genetic correlations are valuable when they allow
Douglas-fir breeding programs emphasize a breeder to use indirect selection. Although stem
productivity which is measured as height and
diameter. Because these are the mostly
frequently measured traits, more is known
about their heritabilities than for other traits.
Heritabilities for growth traits typically range
from 0.10 to 0.30. Because growth traits have
low heritabilities, most breeding programs
rely heavily on among family selections to
obtain genetic gain. Heritabilities of family
means tend to be much higher (0.60–0.90)
because families are typically planted on 4 or
more sites and are usually represented by
more than 60 progeny (Howe et al. 2006). The
added benefit of replicating families over
multiple sites is the possibility to reduce the
impact of genotype by environment
interaction by finding families that perform
well and are stable across a breeding zone
(Stonecypher et al. 1996). Heritabilities must
also be examined in the context of age
because heritabilities for growth traits slowly
increase with age (Johnson et al. 1997).
Genetic correlations
Genetic correlations are important because
breeders may cause undesirable changes in
some traits by selecting for other correlated
traits. In Douglas-fir increased growth is
associated with increased second flushing, late
bud set, and increased cold injury in fall.
These adverse relationships are stronger
among than within populations. In contrast
there is no con- sistent correlation between
growth, cold injury in spring or bud burst
(Howe et al. 2006). Wood density consistently
shows an adverse genetic correlation with
growth. This association is stronger for
diameter than for height growth (King et al.
1988a, El-Kassaby and Park 1990, St. Claire
1994). Howe et al. (2006) cite two studies
(King et al. 1988a, Vargas-Hernandez and
Adams 1992) in which the adverse genetic
correlation was strong, wood density
decreased by 3 to 6.5% by selecting for
increased growth using a selection intensity of
10%. One reason for their modest loss in
wood density is that density has a high
heritability, but a small coefficient of additive
120 Douglas-fir: The Genus Pseudotsuga
volume is the primary trait of interest, selection is 3,456 seedlings were planted. Mortality was highest
usually based on height and diameter at
measurements. These traits are highly correlated,
with genetic cor- relations usually 0.80 or higher
(Yeh and Heaman 1982, Johnson et al. 1997), but
lower correlations between height and diameter
are sometimes found (e.g., 0.45 by King et al.
1988b). The timing of bud set in first-year
seedlings, which have seasonally indeterminate
growth, has a negative genetic cor- relation with
fall frost hardiness, those that set bud early are
more frost hardy. In saplings that have seasonally
determinate growth, and set buds earlier in
summer, the correlation with fall frost hardiness
in weak (Li and Adams 1993). The positive
correla- tion between bud break in spring and
spring frost hardiness is stronger and less
influenced by age. Trees that flush late in spring
are more frost hardy. Correlations between cold
hardiness in spring and fall seem to vary by
population. Such correlations are not found in
seedlings and saplings of a Cascade population or
were weak to moderately negative in a coast
population (O’Neill et al. 2000, 2001).
Estimates of genetic gain
Realized genetic gains may be determined by
com- paring the performance of genetically
improved materials in genetic gain trials. Stoehr et
al. (2010, 2011) established realized genetic gain
trials in British Columbia on five low-elevation
sites representing a range of site indices.
Populations of three types of genetic quality were
chosen as control unimproved (from a mix of wild
stand seed lots); elite (by cross- ing the best nine
parents to yield an average breed- ing value of 18
(that is should produce 18% more volume at
rotation age than nonimproved popula- tions;
intermediate (obtained by mating of parents of
somewhat lower ranking than the top parents
yielding a breeding value of 10.
Seedlings were planted as 1-0 container stock
in
1996 at four different planting spacings: 1.6 x 1.6
m, 2.3 x 2.3 m, 2.9 x 2.9 m, and 4.0 x 4.0 m; these
yielded an equivalent of 3906, 1890, 1189, and
625 trees/ha, respectively. Sites were established
with two replications each of 12 x 12-tree square
plots (144 seedlings) per genetic level and spacing
com- bination. Therefore, on each site a total of
the 1.6 x 1.6-m spacing compared to the three 5.4% for height, and 6.4% for diameter. The
other spacings. The intermediate population results from the realized genetic gain trials in
suffered the highest mortality in all four British Columbia and Oregon demonstrate that
spacings, while elite and control populations progress from selection and breeding of coastal
survived best. Overall mortal- ity, however, Douglas-fir is achievable.
was low. Realized gains at age 12 for the elite
progeny were 48% for volume and 15% for
height, compared to predicted genetic gains of
36% for volume and 18% for height. Realized
gains for the intermediate progeny were 29%
for volume and 10% for height, compared to
predicted gains of 20% for volume and 10%
for height. The realized gains were above
control means across all sites and spacings.
The Northwest Tree Improvement
Cooperative (NWTIC) and the Pacific
Forest Research Station, initiated a realized
genetic gain study along the west slopes of
the northern Oregon Cascades in 1997 (Ye et
al. 2010). Parents were selected from the first-
generation Molalla breeding zone in the
Cascade foothills southeast of Portland,
Oregon. Populations of three kinds of genetic
quality were chosen: elite (created by single
pair matings of 20 top parents); intermediate
(created by single pair matings of 20 parents
of somewhat lower ranking than the top par-
ents); unimproved (the unimproved
population was a random selection of 50 trees
selected from naturally regenerated stands
well distributed throughout the breeding
zone). One or two transplant seedlings of
the elite, intermediate, and unimproved
popula- tions were planted in spring 1997 at
five sites in the Molalla breeding zone. In
each of the six replicates at each site
seedlings of the three genetic quality types
(elite, intermediate, unimproved) were
planted at each of two stand densities: low
density (3.6 x 3.6-m spacing, 772 trees/ha)
and high density (1.8 x 1.8-m spacing,
3,086 trees/ha) using a split-plot design with
planting density as the whole plot and
genetic quality as the subplot. Each split-
plot had 100 trees arranged in a 10 x 10-tree
square.
The realized gains at age 15 averaged over both
the elite and intermediate progeny were
17.2% for stand volume per hectare, 3.5% for
height and 4.5% for diameter compared to
predicted genetic gains of 16% for volume,
Chapter 5. Tree Breeding and Improvement 121

Breeding Goals and Objectives and Joyce 1990, Campbell 1964, Wheat and Silen
The two main objectives of Douglas-fir breeding 1977). Greater growth results in greater yields at
programs in western North America are to harvest or permits shortening of rotations. The
improve economic traits, that is, to achieve key traits used as predictors of rotation age are
increased crop value, and to ensure that the DBH and total height, usually measured anywhere
resulting breeding populations are well adapted at ages 5 to 20. In addition to increasing tree
and have sufficient genetic variation for gains to value, most breeding programs seek to maintain
continue in subsequent generations (Johnson sufficient physiological adaptability and genetic
1998). Although the value of tree crops is variability. In fact, the existence of breeding
determined on a per-hectare basis, the traits used programs to maintain adaptability is a hallmark of
for selection are usually measured on individual Douglas-fir breeding (Howe et al. 2006).
trees because it is difficult and expen- sive to
measure the performance of many families on a Secondary breeding objectives
per-hectare basis. Breeders, however, must be Improvements in wood properties and stem
mindful of the assumption involved in defining quality are secondary in importance because their
breeding objectives on a per-tree basis when the impact on tree volume is neither as great, nor as
real goal is to increase the value of the entire crop quantifi- able, as it is for stem volume (Howe et
(Howe et al. 2006). al. 2006). The single most important wood
property is wood specific gravity because dense
Primary breeding objectives wood is associated with wood strength as well as
The two main breeding goals—increasing crop increased pulp yields. Important stem defects
value and maintaining adaptability—are generally include forks and ramicorn branches. Forked
met in different ways. The most important stems are formed when the ter- minal leader is
adaptive traits are cold and drought hardiness. damaged or killed and two lateral branches
They are usually maintained through the use of subsequently assume equal dominance. Ramicorn
appropriate breed- ing zones. The reason for that branches are excessively large, upright branches.
is three-fold (Howe et al. 2006). First, because Forks and ramicorn branches tend to make a
severe frosts and droughts are rare, it is difficult to portion of the stem unmerchantable. Genotypes
measure frost and drought hardiness under normal with many stem defects are unlikely to be
progeny test conditions. Secondly, how these rare included in future breeding populations.
events will affect produc- tion plantations is Selections for in- creased wood density are often
difficult to predict and thus the value of made only among the fastest-growing genotypes
improving cold and drought hardiness. Thirdly, in a two-stage selection. Few insect and disease
inexpensive artificial tests can be used to measure problems have risen to the level where they form
cold and drought hardiness (cf. chapters on frost key components of Douglas-fir breeding
and drought). Most of the parents in first- programs. The steadily increasing infec- tion of
generation breeding populations were selected Douglas-fir stands in coastal areas of Oregon,
from the breeding zones in which they will be Washington, and British Columbia with the
used; thus damage from cold and drought should pathogen Phaeocryptopus gaeumannii in the last
be compa- rable to that in natural populations. decades of the 20th century has become a matter
In contrast to adaptability, tree value is gener- of great concern. Hence, tolerance to Swiss needle
ally improved by selecting and breeding the most cast has become a breeding objective. McDermott
valuable genotypes within these well-adapted and Robinson (1989) demonstrated significant
pop- ulations. Tree value is primarily determined variation in resis- tance to the pathogen among
by stem volume and secondarily by stem quality, nine provenances from California, Oregon,
and by wood properties such as density (Howe et Washington, and British Columbia. Studies
al. 2006). Therefore, the primary breeding carried out within the framework of the Swiss
objective for Douglas-fir is to increase volume Needle Cast Cooperative suggest that tolerance to
growth (Adams Swiss needle cast can be improved via
selection and breeding (Johnson 2002).
122 Douglas-fir: The Genus Pseudotsuga

Steps in Tree Breeding Programs


Genetic Improvement Programs
Howe et al. (2006) listed four main steps for a
typical tree-breeding program:
North America
The first step is to delineate breeding zones. The Woods (1993) and Lipow et al. (2003) described
second step is to develop one or more breeding major Douglas-fir genetic improvement programs
populations for each breeding zone. In the first in North America. Their accounts were updated in
generation, breeding populations may be selected
from wild stands within each breeding zone, from 2006 by Howe et al. Most Douglas-fir
superior non-local popula- tions that have been improvement in North America is carried out by
identified based on provenance test, or from land the Northwest Tree Improvement Cooperative
races. The first approach is common for Douglas-fir
in its native range, whereas the second approach has (NWTIC), Inland Empire Tree Improvement
been used for Douglas-fir in Europe and the southern Cooperative (IETIC), British Columbia Ministry
hemisphere. The third step is to field test the progeny
of Forests (BCMoF), and Weyerhaeuser
of selected parents and pursue advanced- generation
breeding within each breeding population. The fourth Company. These four organizations are
step is to produce genetically improved materials for responsible for developing improved materials
outplanting. In most Douglas-fir programs, this
planted by private companies, tribal governments,
involves establishing the best genotypes in separate
wind-pollinated seed orchards. and public agencies in the USA and Canada.

Breeding zones Northwest Tree Improvement Program (NWTIC)


Breeding zones are used to manage the Douglas-fir tree improvement programs in the
deployment of trees from breeding programs and Pacific Northwest began in the 1950s when a
wild seed col- lections, respectively. A breeding small group of government agencies and timber
zone is a group of sites across which a breeding companies began to select coastal Douglas-firs
population can be planted and expected to and established clonal seed orchards. The
perform well. Two vastly different approaches Industrial Forestry Association (IFA) was
have been used to delineate breeding zones: 1. instrumental in these undertakings by hir- ing
direct approaches based on long- term field tests John Duffield, a forest geneticist, to guide their
of breeding materials, and 2. indi- rect approaches tree improvement efforts (Hagenstein 1966). Roy
based on seedling tests of natural populations. A Silen, a USDA Forest Service geneticist, proposed
major disadvantage of the indirect approach is the in 1966 a “progressive tree improvement
unknown relationship between the genetic program” based on results from the 1912
distance measured in indirect tests and the Douglas-fir hered- ity study (Silen 1966a). His
genotype by environment interactions measured proposal was accepted the same year by the IFA.
in long-term field tests (Howe et al. 2006). The name progressive infers incremental genetic
gains with each succeed- ing seed crop as
Selection information about parent trees improves (Silen
Because of the long rotations in forest trees, selec- and Wheat 1979). The Progressive Tree
tions must be made long before harvest age. In Improvement Program was implemented by
Douglas-fir, final selections are commonly made forming local, geographically based cooperatives
when trees are about 10 to 15 years old (Howe et to share the costs and benefits of tree
al. 2006). Based on age-age correlations from improvement. The IFA Progressive Tree
more than 51 progeny test sites in Oregon, Improvement Program evolved into the Northwest
Johnson et al. (1997) concluded that per-year Tree Improvement Cooperative (NWT1C) in
gains are maximized when selection are made for 1986. As of 2004, the NWTIC, housed at Oregon
height at age 10, and for diameter at age 13. State University, consisted of 27 member
For improving tolerance of Douglas-fir to organizations, representing forest industries, tribal
Swiss needle cast disease, early selection in the governments, state and federal agencies within the
field at age 2 years was 25% to 100% as efficient USA and Canada (Howe et al. 2006). Distinctive
as waiting until age 10 or 12 years (Temel et al. features of the Progressive Program included the
2004, 2005). assumption that local seed sources are best in the
mountainous and environmentally
heterogeneous Douglas-fir region, low-
intensity selection of first-
Chapter 5. Tree Breeding and Improvement 123
the Canadian Forest
generation parents, use of many small breeding
zones, and the use of very large breeding
popula- tions. This approach was intended to
ensure adapt- ability of the first-generation
breeding populations, which consisted of parents
selected from natural stands within the breeding
zone (Silen and Wheat 1979). Although breeding
zones have been consoli- dated and breeding
populations are being reduced, the NWTIC
program is still one of the largest tree breeding
programs in the world (Howe et al. 2006). When
the establishment of first-generation tests was
completed in 1993, 21 first-generation coopera-
tives had been formed and the Douglas-fir zone
west of the Cascades was blanketed with 109
breeding zones ranging from the Canadian
border to north- west California. More than
26,000 first-generation parents have been
evaluated based on more than three million
progeny test trees (Lipow et al. 2003). The large
number of breeding zones was recently reduced
to eight second-generation zones and the
number of parents used in advanced-generation
breeding is expected to be about 2,000 (Howe et
al. 2006).
Weyerhaeuser Company
Weyerhaeuser Company has managed a tree
breed- ing program since 1963. One of their key
assumptions was that rigorous phenotypic
selection of superior trees in natural stands would
produce genetic gain in growth. Therefore, the
foundation of their pro- gram was an intensive
plus-tree selection program in stands 25- to 80
years old. Some 3500 parents were se- lected in
six breeding zones covering Weyerhaeuser lands
in Washington and Oregon (Stonecypher et al.
1996). The primary objective of the first-
generation program was to improve growth and
stem quality. Selection, breeding, and testing of a
large second- generation population are almost
complete, and the third-generation of
improvement is underway (Howe et al. 2006).
British Columbia Ministry of Forests (BCMP)
In Canada, first-generation testing began about
I960 for coastal Douglas-fir and 1980 for interior
Douglas- fir. Historically, tree improvement
programs were coordinated by cooperative tree
improvement coun- cils, consisting of the BCMF,
methods of seed production, and conservation of
Service, private companies, and universities. ex situ ge- netic resources of Douglas-fir in
A Plus Tree Board was formed in the 1960s, Europe. In 1986, the Federal Republic of
followed by the Coastal Tree Improvement Germany had 1,346 Douglas-fir
Council in 1979, and the Interior Tree
improvement Council in 1981. These two
councils were later merged into the Forest
Genetics Council of British Columbia (FGC)
in 1998. In the coastal program, half diallel,
factorial, and open-pollinated mating designs
have been used to test about 660 parents in
130 field tests. In the in- terior program, an
open-pollinated mating design was used to
test about 1,600 parents in 32 field tests
(Howe et al. 2006).
Inland Empire Tree Improvement Program (IETIC)
The IETIC was formed in 1968 to develop
improved ponderosa pine for the Inland
Empire, a region including eastern
Washington, eastern Oregon, northern Idaho,
and western Montana. The IETIC is housed at
the University of Idaho and consists of 19
organizations, including federal and state
agencies, private companies, tribal
governments, and universi- ties. A Douglas-fir
species group was formed in 1974. In most of
the 13 breeding zones, 200 to 300 trees were
selected, and more than 2,500 first-generation
parents have been field-tested to date.
Compared to the Pacific Northwest, Douglas-
fir is relatively less important in the Inland
Empire. The area planted to Douglas-fir has
declined in recent years because reliance on
natural regeneration rather than on plant- ing
is now common. Consequently, the Douglas-
fir breeding program in the Inland Empire is
much less intensive than it is in the Pacific
Northwest (Howe et al. 2006).
Europe
Interest in Douglas-fir improvement has
grown in conjunction with the growing use of
Douglas-fir in European forestry.
Collaboration among forest re- search
organizations in several European countries
has led to the formation of a European
Douglas-fir Improvement Research
Cooperative (EUDIREC). Among its
objectives are the development of a data- base
for European gene resources, the building of a
common breeding population, improving
124 Douglas-fir: The Genus Pseudotsuga
stands covering 1,453 ha, certified as selected planting of Douglas-fir began in New Zealand
seed sources for Douglas-fir representing ex situ in the early 20th century. Large provenance trials
genetic resources of Douglas-fir in Germany were established in the mid-20th century, with
(Ruetz et al. 1990). The number of such stands prove- nances from seed collections in North
has probably grown since the reunification of America and from existing Douglas-fir stands
Germany. in New Zealand (Ruetz et al. 1990; personal
communication, HP Schmitt and W Ruetz,
New Zealand 1990). A Douglas-fir breeding program was
Douglas-fir is second only to Monterey pine as a begun in 1970 by M.D. Wilcox of the New
plantation species in New Zealand. Large-scale Zealand Forest Research Institute.
6. Flowering
Denis P. Lavender

T
he initiation of reproductive structures in struc- tures of conifers, the role of juvenility and
plants has been the subject of many studies maturity
(Evans 1971; Zeevaart 1976; Wareing 1980; Bernier
1981, 1988; Bernier et al. 1993). Most of these
studies, however, have been about herbaceous
annual plants. The reproductive phase of these
plants differs sig- nificantly from that of woody,
perennial plants in that (1) upon receipt of an
external signal (such as a critical photoperiod), the
annual plant can convert all apical meristems on
the shoot to flower primordia, whereas the
perennial plant must conserve many such
meristems for future vegetative growth; and that
(2) the mature perennial plant has what is called a
“burden of history” that can affect the incidence
of flower production in any given year.
Previous reviews of the literature on flowering
in angiospermous forest trees (Lavender 1984),
for tem- perate zone woody plants (Jackson and
Sweet 1972), and for temperate zone coniferous
trees (Lavender and Zaerr 1985, Puritch 1972)
demonstrated that relatively little research effort
has been focused on flowering in large, woody
perennial plants; that flowering may be stimulated
by a range of cultural treatments; and that
application of plant growth regulators may
stimulate differentiation of reproduc- tive
structures. Evans (1971) noted that more than
1,200 papers on flowering had appeared in a 6-
year period, and there have been many more since
then. Although most were concerned with
herbaceous angiosperms, enough discussed
conifers and were sufficiently repetitious;
therefore, it is most practical if we limit much of
our discussion to material in the many review
papers.
Three major topics important to flowering are
the
history and nomenclature of the reproductive
in flowering, and the incidence of flowering and definition of a flower as, “a determinate
methods of eliminating the initiation of flowering. sporogenous shoot,” which qualifies coniferous
reproductive structures as “flowers.” Actually, the
History and Nomenclature staminate strobilus (male flower) consists of
A flower has been defined as “a determinate several spirally arranged microsporophylls, each
axis with spore-bearing appendages (and, of which bears two microsporangia on the lower
usually, sterile appendages) and short surface. The ovulate strobilus (female flower)
internodes occurring in the angiosperms” comprises a central axis that bears ovulate scales,
(Bold et al. 1980, p. 748). Douglas-fir each borne in the axis of a bract. Each scale has
reproductive structures, in common with two ovules on the upper surface, each of which
those of other conifers, are frequently called consists of an integu- ment surrounding a
“flowers,” but they lack the calyx, corolla, megasporangium (Figure 6.1). There has been
stamens, and pistil com- mon to the flowers some debate over the morphol- ogy of female
of angiosperms, and they cer- tainly do not reproductive shoots. Florin’s classic monograph
qualify under the definition of Bold et al. (1954) noted that the first three decades of the
(1980). They also do not meet the definition 20th century were dominated by discussions of
of “a determinate sporogenous shoot that the “Sachs-Eicher excrescence” (female of coni-
bears carpels,” as offered by Romberger and fers as a simple flower) and brachybest (megaspo-
Gregory (1974, p. 138). Jackson and Sweet rophylls) florescences theories (pp. 380–384). By
(1972), however, proposed a simplified 1940, the brachybest theory was favored, although

125
126 Douglas-fir: The Genus Pseudotsuga

Seed Pollen
1
Bud Swelling
Bud tips show lighter color
Duration: 10-20 days

2
Bud Bursting
1/4 of strobilus protrudes
through ruptred bud
scales Duration: 3-5
days

3
Strobilus Emerging
1/4-1/2 of strobilus
exposed; Duration: 2-4
days

4
Strobilus Emergent
1/2-3/4 of strobilus exposed;
bracts not fully reflexed;
pollen sacs separated and
dry Duration: 2-3 days
5
Pollination
>3/4 of strobilus exposed;
bracts fully reflexed;
pollen sacs shedding
Duration: 2-3 days

6
Pollination Past
Cone scales visible between bracts
and appressed; pollen shedding completed,
Pollination completed 10-17 days
following onset of bud bursting

7
Seed One Pendant
17-24 days after bud bursting

Figure 6.1 Seed and pollen development.


Chapter 6. Flowering 127
According to Brink (1962):
Florin argued that more research was necessary to
precisely define the coniferous cone. He observed
that the female “flower” of conifers has been of
interest since 1950, and that early research on
flow- ering was largely confined to the
classification of megasporophylls. Florin also
noted that “the taxads are accordingly regarded as
forming a class of their own—Taxopsida—
distinctly separate from the co- nifer opposites”
(Florin 1954, p. 382). Conifers have and taxads
have been considered an ancient and relic group
and have been referred to as “living fossils”
(Williams 2009, p. 3). Several researchers have
listed both conifers and taxads under the order
Coniferophyta (Gifford and Foster 1988, pp. 402–
404; Bold et al. 1980, pp. 506–508).

The role of juvenility and maturity in


flowering
Wareing’s (1959) review of the literature
concluded that juvenility and maturation each
represent two stable states of the plant, and that
aging describes the transitory states during the
increase in size as the plant grows. Further,
several morphological fea- tures change gradually
with the change to maturity. The plant’s capability
to flower is perhaps the most notable. In his
description of the general growth curve of a
perennial plant, Wareing (1958) observed that the
production of female flowers begins on the
ascending limb of the curve and that male flowers
are generally initiated on the descending limb.
This relation between flowering and the reduced
vigor holds not only for the whole plant, but for
individual branches. Other important points are as
follows:
• Plant size and duration of the juvenile period
are roughly correlated.
• Plants in the juvenile stage generally root
readily, but mature plants commonly do
not.
• Nutrition is generally higher in juvenile than
mature plants.
• The phase change between juvenile and
mature plants is associated with size and
complexity of the shoot and not with the
number of annual cycles it has passed through.
• Both “juvenile” and “mature” status are stable.
allowed.
Phase change sometimes involves a relatively
abrupt switch in potential of apical meristems in Romberger (1967) believed that the knowledge of
higher plants from a juvenile to an adult type of the physiology and development of flowering in
growth. The two growth forms are highly
persistent in different parts of the same
trees is inadequate and therefore difficult, and that
individual and in clonally propagated offspring. explain- ing contradictory results is also difficult:
Reversion to the juvenile condition occasionally “The large
occurs in buds on adult-type shoots, and
invariably takes place in sexual reproduction. It
is pointed out that counterparts of phase change,
expressed in a wide variety of ways, are common
among both plants and animals; structures arise
in the development of all organisms with charac-
teristics that are not merely functional
adaptations but are innate and self-maintaining in
mitosis. Phase change is to be considered,
therefore, as illustrating a general aspect of cell
heredity and variation; it is singular only in the
distinctness with which the alteration in type of
growth may be phenotypically expressed. The
mecha- nism whereby such characteristics are
maintained and perpetuated in somatic cells is an
unsolved problem. There is now limited
evidence, some of it also indirect, suggesting that
the chromosomes are the site of such
discontinuous and potentially reversible onto-
genetic changes. This evidence is briefly
reviewed in terms of a hypothesis in which it is
assumed that in addition to the genes, which are
stable, the chromosomes also contain self-
perpetuating accessory materials that un- dergo
paramutation in an orderly way in somatic cells
as an essential aspect of a nucleo-cytoplasmic
system of morphogenetic determination. (Brink
1962, p. 1)

Romberger (1967, p. 2) asks:


What is the state of our present knowledge of the
control of flowering trees? In a very general way
we can say that even on good sites young trees
do not flower until they have attained a certain
minimum size, complexity of branching, and
“physiological age.” Such a statement is not an
explanation, but only an enumeration of condi-
tions which usually, or sometimes, are
prerequisites to the onset of flowering. It does
not address itself to the problem of the means by
which it is determined that a bud shall develop
into a flower rather than a shoot. The
physiologist who has attempted to analyze the
problem of flowering in trees is apt to find
unconvincing any general statement about its
control. This is because of the diversity, or even
contradiction, of actual observations and the
paucity of data from well-designed experiments.
Discussion of the problem of flowering with
reference to the literature is difficult without
involving concepts of “juvenility,” “ripeness to
flower,” “maturity,” or “physiological aging.”
These concepts are not amenable to precise
definition, and they cannot be given much
concrete content until our understanding
penetrates much deeper than it yet has.
Nevertheless, for the pres- ent, they are useful in
spite of their lack of precision because they
delineate comprehensible areas of concern from
the vastness of developmental biology. Similar
use of a term such as “epigenetic” should also be
128 Douglas-fir: The Genus Pseudotsuga
and poorly defined area including juvenility, Attainment and maintenance of the ability or potential to
flower is the only consistent criterion available to assess
aging, apical dominance, gravimorphism, and the termination of the juvenile period. Other characteristics
growth habit also needs additional attention. known to change with development and/
Failure to flower is usually concurrent with the
juvenile form, but the two may really be
independent (p. 24).”
The concept of juvenility has been questioned,
in that all attributes that are said to be a function
of it are not expressed simultaneously, and that it
does not fulfill the requirements of a systemic
analysis (Borchert 1976). Therefore, as Borchert
argues, “the existence of one uniform juvenile
state must be se- riously doubted based on the
available evidence” (Borchert 1976, p. 22).
Sussex (1956, p. 271) asked two questions: “Does
the plant body attain some critical size and then
signal the meristem to initiate the developmental
phase-change response, or does the meristem
behave independently of the remain- der of the
plant?” His review of papers by Wareing answer
the first question affirmatively.
Schwabe (1976) noted that the juvenile stage
may have an important role in a plant’s survival,
in that energy required for height growth will
not be di- verted to flowering until the plant has
successfully competed with neighbors. Citing
results of experi- ments with several plants,
including trees, Schwabe hypothesized that
flowering may be controlled by the distance a
plant apex is from the roots or by root activity,
and suggested that gibberellins (GAs) are the
likely cause. Further, he noted that root disease
frequently causes cones to form in Japanese
larch and Sitka spruce (as it does in Douglas-
fir).
For woody angiosperms plants, Wareing and
Phillips (1970, p. 61) suggested that, “low
gibberel- lin levels are a necessary but not
sufficient condition for juvenile/adult transition.”
Ross (1976) reported flowering on Douglas-fir
grafts of mature tissue, but not on equal-sized
seedlings, concluding that the age of the
meristem, not plant size, controlled the juvenile-
to-adult phase change in this species. Longman
(1976) argued that juvenile trees may be forced to
flower and that basal girdling is a power- ful tool
for doing so.
In reviewing the literature on juvenility and
ma-
turity, Hackett (1985) noted:
or age are not consistent from species to important in promoting or inhibiting phase-change or
species, and none has been demonstrated to be its reversal, but phase-change is an inductive
causally related to sexual maturity . . . it is phenomenon and once it has taken place the
possible to show that there is a transitional conditions that promoted the change need not
phase of development during which flowering continue to operate. (Wareing 1987, p. 90)
potential is increasing. (Hackett 1985, p. 111)

Plants grown in a low rainfall area of


southern Vancouver Island started flowering
at 6 to 10 years of age, whereas those grown
50 miles away in a cool, wet region started
later. Pharis et al. (1980) concluded that
“the best way to shorten the juvenile period
is to grow the plant rapidly to a minimum
size and then apply a flower inducing
treatment, both appropriate for the given
species”:
It is now not known why attainment of a
minimum size is required for transition to the
mature condition. The observations that the
juvenile-to-mature phase change usually occurs
at a predictable stage (size) in the develop- ment
cycle of a plant and that changes occur at the
shoot apical meristem raise questions about the
organismal locus of the phase change: Does the
plant body attain some critical size and then
signal the meristem to initiate the developmental
phase change response? Or does the meristem
behave independently of the remainder of the
plant? (Pharis et al. 1980, p. 120)

In one of his last papers, Wareing (1987)


asserted the following:
The properties of the apex are determined by the
struc- ture and organization of the apex as a
whole, and the differences between the two
phases (juvenile and adult) arise from the
different organization of their apices. . . .
There are intrinsic differences in the
meristematic cells of the juvenile and adult
apices, and the differences in their properties and
structure arise from intrinsic differences in the
cells. It is suggested that phase change
indicates
the existence between cells with respect to gene
expres- sion, which can be transmitted through
repeated cycles of cell division. Persistent
differences in gene expression, without
permanent changes in the genome, are said to be
epigenetic. (Wareing 1987, pp. 85–86)

This hypothesis is supported by the fact that


the DNA of apical meristems of juvenile and
adult shoots in Hedera helix (English ivy, a
plant frequently used to study phase transition)
is the same. Finally, Wareing (1987) wrote,
It has been argued that phase change must
involve in- trinsic differences in juvenile and
adult apical meristems, but this does not exclude
the possibility of influences from the mature
parts of the plant, especially hormonal factors
arising in the leaves or roots. In particular, influ-
ences from other parts of the plant may be
Chapter 6. Flowering 129

Poethig (1990) summarized the role of phase as maturation processes, the vegetative
and phase change in detail, using primarily propagation of woody plants becomes
genetic data: increasingly difficult” (p. 19). This aspect of
• Six gene loci individually or maturation has been of inter- est in Douglas-fir. It
collectively govern the transition from will be discussed further in the section on rooting.
vegetative to reproductive state. According to Greenwood (1992),
• Several anatomical and morphological states The onset of the mature state is usually gradual. For
example, grafting studies on loblolly pine and eastern
include foliage characteristics, phyllotaxy,
larch have shown that maturational change is gradual
plastochon, growth habit, aerial root rooting for most traits, except for plagiotropism and branch
ability in much studied Hedera helix frequency. But plantlets derived from buds induced on
associated with juvenile and adult states that the cotyledons of mature embryos from several
are regulated independently of flowering. conifers exhibit mature characteristics immediately
after they begin to grow . . .
• The three phases are juvenile vegetative, Maturation affects a wide variety of
adult vegetative, and an adult reproductive. morphological, physiological, and biochemical traits,
but these traits ap- pear to vary independently of one
• Phase changes regulated by factors both
another. Maturational traits are often persistent, and
intrinsic and extrinsic to the terminal meristem. their maintenance is not always a function of tree size
• The compound florigen’s origin has not been or proximity to roots. Differences in chlorophyll
content, specific leaf weight, and xylem morphology
isolated and probably will not be in the near
among eastern larch scions from trees of different
future. ages grafted at the same time persist for several years.
• Factors that increase growth generally tend to On the other hand, anecdotal evidence suggests that
mature morphological characteristics disap- pear
promote flowering; factors that reduce growth
quickly after grafting or rooting cuttings in Populus
inhibit flowering. spp. and Eucalyptus spp.
• The whole concept of phase change represents Evidence suggests that the cells of the apical meri-
an extremely complicated growth pattern. stem itself become determined in some woody plants.
Grafted apices from mature plants of Citrus spp. or
In summary, “genetic, developmental, and Sequoia spp. consisting of only the apical meristem,
molecular analyses of mutations that affect the and one or two leaf primordial, grow out into plants
with mature characteristics. (Greenwood 1992, p. 21)
expression of particular phases of shoot
development are begin- ning to yield a clearer Greenwood and Hutchinson (1993) observed
picture of the regulatory framework of shoot that four major methods to maintain juvenility
development” (Poethig 1990, p. 929). have been frequently studied: serial propagation,
Hackett and Murray (1992) suggested that the hedging, repeated subculture, and tissue storage.
maturation process (phase transition) in woody Hedging has failed with Douglas-fir. They also
plants is “a very dramatic and protracted example argued that reproductive competence is not a good
of determination and differentiation” of an array measure of maturity. Studies with several species
of phenotypic characteristics that may not be strongly suggest that male:female ratio is a better
closely linked temporally or mechanistically (p. measure of maturity than is the presence of
197). Hackett et al. (1992) proposed that phase strobili alone. (This measure is true with Douglas-
development, “is not fundamentally different from fir, according to numerous observations by
plant develop- ment in general but may have Lavender: young seed- lings produce more female
unique features that are particularly important to flowers than male.) In concurrence with other
understand,” and that “phase-related changes in researchers, Greenwood and Hutchinson (1993)
phenotype are the result of subtle changes in the argued that minimum size and minimum maturity
gene expression that overlay the fundamental are necessary for flowering; however,
patterns of gene expression that is common to “experiments with grafted scions do not, in our
both the juvenile and a mature plant body” opinion, indicate that the ability to flower is solely
(Hackett et al. 1992, pp. 84–85). a function of increased size and complexity of the
Greenwood (1992) noted that maturation “has re- plant” (p. 24). They noted that information was
ceived renewed attention over the last 10 years lacking on “the relationship of gene expression to
since, phase change in woody plants. Therefore, models
130 Douglas-fir: The Genus Pseudotsuga
for the role and the regulation of gene expression as-
in maturity are more speculative than predictive.”
Additionally, some data suggest that, “genes play
a role in the regulation of maturity,” and that
control of any one trail of maturity is
independent of that for others (Greenwood and
Hutchinson 1993, p. 26). In a later paper on
conifer juvenility and matu-
ration, Greenwood (1995) offers this explanation:
Maturation is an integral part of the life cycle of all
vascular plants. Four phases of maturation have been
recognized: (1) the embryonic phase, (2) the post-
embry- onic juvenile vegetative phase, (3) the adult
vegetative phase, and (4) the adult reproductive phase.
In woody plants, maturation is associated with
decreased growth rates during phases 2-4 (which
often persist in vegeta- tive propagules), increased
plagiotropism, and changes in reproductive
competence, branching characteristics and foliar
morphology. In addition to morphological changes,
there are numerous physiological and bio- chemical
changes that accompany the transition to the adult
phase. . . .
Maturation involves changes in the habitual
behav- ior of meristems, where habit is defined as a
behavior pattern that develops in response to a
particular set of physiological inputs. By definition,
habits tend to resist change, but nonetheless can be
altered in response to varying environmental inputs.
The earliest maturation event is the polarization of the
embryo into roots and shoots, where each meristem,
starting with the same genes, acquires unique habits
adapted to different envi- ronments. In shoots, the
meristems continue to develop new habits as the plant
grows. Over time, the meristems of conifers change
their behavior, losing regenerative potential and
capacity for vegetative growth, but gain- ing
reproductive competence and more massive leaves.
(Greenwood 1995, p. 493)

The above survey, although admittedly brief,


refers to most of the reports relating the juvenile
and mature phases in woody coniferous plants to
the incidence of flowering in these trees. None of
the reports was concerned with Douglas-fir,
although studies by Owens (1984a,b) related the
effects of treatments that stimulate flowering on
the anatomy of apical and lateral meristems with
that of control trees (see also Owens and Blake
1985). Because the trees were all probably
juvenile (10 years old), the data could not be
interpreted in terms of potential differences
between juvenile and mature shoots. Although
some disagreements were described by Owens
(1984a,b), the general conclusion remains the
same as that found in an early review (Mathews
1963), which noted that flowering in conifers is
sociated with the adult state. Unfortunately, bud on the tree. For floral development to be
while subsequent work qualifies as basic ‘evoked’ or initiated, all systems must be
research, care- fully done, it does not identify ‘permissive,’ if any one system is ‘non-
the basic cause of the differentiation change permissive,’ vegetative growth prevails.”
resulting in flowering.
Romberger (1967) argued that flowering
is a de- velopmental phenomenon—a point
with which later researchers agreed
(McDaniel et al. 1987)—and concluded that
“strenuous efforts should, I think, be made to
collect, organize, and evaluate present in-
formation on development physiology in trees
with full references to related information
and thinking in other areas of biology” (p.
11).
Incidence of flowering
The work described above correlates the
incidence of flowering strongly with the
development of the adult phase, not
surprisingly because the ability to flower often
defines maturity, but does not identify the
biochemical reasons for either. Similarly,
several reports have noted that flowering in
conifers is very erratic (Matthews 1963).
Romberger and Gregory (1974) observed that
“the subject of flowering in trees” is “vast,
complex, and confusing. A subject of such
breadth and importance needs a firm base of
good detailed, descriptive, developmental and
experimental studies. That base is now quite
inad- equate and it is not being enlarged very
rapidly” (p. 132). For example, among 4,073
papers about trees published from January
1970 to June 1974, none were about
flowering. As Romberger and Gregory (1974)
explained, trees are inconvenient subjects for
research. They differ from annual plants, in
that flow- ering is not subject to one or several
environmental factors because trees must
conserve some buds for vegetative growth.
Experiments on flowering focus on single
observation dates, whereas trees have the
burden of years of exposure to environments
that may condition their response to current
trials. Finally, trees have an extended
developmental time scale: months may
intervene between events that stimulate
flowering and the evidence of it. According to
Romberger and Gregory (1974), “numerous
bio- chemical or physiological systems are
involved in the control of flowering in every
Chapter 6. Flowering 131

In a previous paper, Romberger (1967) noted between moisture stress and flowering in trees is
that, “flowering research to date is not capable of inconsistent. Reasons for this may be that
defining the basic cause of flower initiation. moisture stress was frequently not measured and it
Romberger and Gregory (1974, p. 145) state, may often have been too low to promote flower
“Analytical morpho- genesis offers a means of bud formation; Owens and Blake (1985) suggest
attaining a new level of understanding of the that pre-dawn moisture stress of at least 12 to 20
control of flowering in trees.” Unfortunately, little bars is necessary to promote flowers. A drought
in the literature suggests that this procedure has period immediately before shoot formation may
been followed—instead, several papers propose stimulate flowering, as Bonnet-Masimbert and
environmental factors, none of which allude to the Lanares (1978) suggest in a report describing
basic cause of reproductive tissues. flowers on lammas shoots of Douglas-fir.
Sweet (1975, p. 72) noted that both the time Additionally, however, excess moisture (periodic
of flower bud initiation and the developmental root flooding) has also been linked to en- hanced
pe- riod during which the bud becomes flowering in 6-year-old Douglas-fir clones
irrevocably programmed to produce a flower (Bonnet-Masimbert and Zaerr 1987). Although
are important in studies of factors influencing the data to date do not clearly define the role of
cone production. He cites evidence that such a mois- ture in flower initiation, one tenable
period may be lengthy in Douglas-fir. hypothesis is that its effect is mediated through
root physiology. Lavender and Zaerr (1985)
Methodology of flower induction reviewed several pa- pers that discussed the
Despite any remaining gaps in understanding, weak correlation between moisture stress and
there are nevertheless significant contributions cone production for a range of species, but they
in the literature that describe the effects of indicated that evidence for a moisture-related
several major techniques on the initiation of cone crop is not clear. They did note that many
flowering in coni- fers including Douglas-fir. seed orchards in the Pacific Northwest are sited
The following material outlines cultural in areas with pronounced summer drought.
techniques and environmental exposures Puritch (1972) found that moisture supply is
which, singly or in combination, have resulted corre- lated with cone production, but that the
in increased flowering in many trees, in- response to moisture stress is variable and
cluding Douglas-fir. depends on both the species of tree and the time
of year. Moisture stresses of –15 to –33 atm. have
Moisture stress been shown to be effective in stimulating
In a review of factors affecting seed production in flowering in Douglas-fir. Researchers for
trees, Matthews (1963) noted that moisture stress Weyerhaeuser Company, work- ing with
during the summer has long been associated with measurements of internal moisture stress, reported
the production of flower buds. And, in western correlations between moisture stress and
North America, several seed orchards have been increased flowering of conifers: where applied
in the rain shadow of the Olympic Mountains mois- ture stress had no positive effect on
(Lavender and Zaerr 1985). Similarly, Jackson flowering, the treatment failed to produce
and Sweet (1972) observed that flowering is often increased moisture stress on the tree (see
reported to be as- sociated with moisture stress, Schmidtling 1974). Owens (1991) re- viewed
but they cautioned that timing may be important. several references that reported a correlation
Ebell (1967, 1970) correlated summer moisture between moisture stress and flowering in Douglas-
stress with flowering Douglas-fir seedlings and fir and pines, but noted that much of this work did
suggested that the effect is mediated by high not
arginine levels. Ross (1988) noted that drying include the time of floral bud initiation.
seedlings to a foliage moisture content of –2.0 MP Light intensity
(with intervening irrigation periods) stimulated
Higher levels of light intensities generally favor
flowering. But in their review, Owens and Blake
formation of female flower buds in trees and
(1985) argued that, at best, the relationship
lower light intensities may increase male bud
formation, according to Matthews (1963) and
Owens and Blake
132 Douglas-fir: The Genus Pseudotsuga
(1985). Silen (1973) found that shading devices fer- tilizer on conifer cone production. An early trial
in- creased flowers of both sexes, but his shading of effects of nitrogen on Douglas-fir is described by
devices raised the temperatures; thus, his results Stoate et al. (1961). They found that nitrate nitrogen
may have been mediated by temperature rather
than light intensity alone. In their review, Bonnet-
Masimbert and Zaerr (1987) cite unpublished
material showing a strongly promotive effect of
high light intensities on Douglas-fir flower
initiation. Owens (1990) sug- gested that light
intensity effects are indirect and related to the
environment within the crown, propos- ing that
the stimulus was perhaps from increased
temperature associated with high light intensities,
rather than the light itself. Finally, Masimbert and
Zaerr (1987) reported that, “increased light
intensity strongly stimulates both male and female
flowering of Douglas-fir” (p. 19); and Winjum
and Johnson (1964) noted that high light favors
reproductive development in Douglas-fir.

Photoperiod
Although the fact that photoperiod affects the
flow- ering of many annual plants is well
known, little evidence of such control has been
shown for woody perennials. One major reason
for this difference is that annual plants can
afford to convert all of their meristems from
vegetative to reproductive uses, but perennial
plants must reserve some meristems for
vegetative growth the following year. Owens
and Blake (1985) suggested that species (such as
those in the Cupressaceae that form male and
female buds at different times) may have the
sex of their flowers determined by photoperiod.
But there is no evidence of such an effect in
Douglas-fir, which initiates both sexes
simultaneously. Certainly the data of Bonnet-
Masimbert (1978), which describes flowers
initiated in the spring and other flowers on
lammas shoots in late summer, are good
evidence of the day-neutral nature of this
species. Owens (1991) and Kozlowski et al.
(1995) both noted that no evidence has been
found indicating that photoperiod-related
flower- ing exists in trees, as it does in
herbaceous plants.

Fertilizers (mineral nutrition)


Numerous publications deal with the effects of
was effective in the production of large cone flowering. Some evidence suggests that natural
crops when applied in the fall. And the same flowering is greater on fertile sites for several
material stimulated more cone buds, both species and that enhanced nutrient
male and female, if applied at time of
vegetative bud break. Ebell and McMullan
(1970) and Ebell (1972) agreed with Stoate et
al. (1961) in that nitrate nitrogen applied at
bud break stimulated the greatest cone
production. In the 1970 paper, Ebell and
McMullan showed that nitrate N stimulated
an increase in amino acids, but the size of the
total free amino acid pool was not related to
cone production. However, levels of the
amino acid arginine and guanidine substances
were raised more by nitrate and were
associated with seed cone production. With
the nitrate ap- plication, ammonia raised the
soluble amino acid content, but not cone
production. Trials on medium and productive
sites resulted in greater responses than trials
on a poor site. Ebell (1972) showed that the
effect of added nitrate was greatest in years
with a good crop of cones on the control trees
and that the strongest responses were when a
dry period followed fertilization. He
hypothesizes (p. 636) that, “reproductive
development is possibly enhanced by an
abrupt properly timed change in nitrogen
status,” and noted that application of more
than 897/kg/ha can lead to toxicity symptoms.
Irrigation can enhance the fertilizer effect, and
a wide variety of pines have shown increased
cone yields with fertilizer treatment under
certain con- ditions. Thus, careful fertilizer
application can ap- parently be expected to
improve cone yield when soil fertility is the
limiting factor in tree growth and vigor, which
is true for most forest sites. Ebell (1972)
noted that the strongest responses to fertilizer
were followed by a dry period.
Several reviews (Owens 1991, Lavender
and Zaerr 1985, Jackson and Sweet 1972,
Owens and Blake 1985, Ross and Pharis
1985, Puritch 1972, and Matthews 1963)
examined many of the reports discussing
mineral nutrition and flower production in
conifers. They agreed that the results of
studies are erratic because time of application,
site qual- ity, and precipitation can all affect
results. To date, we still do not understand the
basic reason why nitrogen may stimulate
Chapter 6. Flowering 133
nitrogenous compounds, especially other guanidine
levels, especially N, are necessary for the growth substances, were higher
of reproductive structures.
Puritch (1972) reviewed several papers
present- ing varied effects of mineral nutrition on
flowering. He noted that Douglas-fir cone crops
have been increased by applying nitrate nitrogen,
but warns that a range of environmental factors
and timing of applications may produce varying
results. He stated, however, that fertilizer
applications are the most common treatment to
stimulate flowering.
Investigation of the effect of mineral nutrition
on flowering in Douglas-fir have been concerned
primarily with nitrogen, although Ebell has
reported that neither phosphorous nor potassium
stimulated flowering. He also noted that
application if nitrate nitrogen (but not ammoniacal
nitrogen) at bud break greatly increased
production of flowers the follow- ing year on 20-
year-old Douglas-fir trees. Though the treatment
did not stimulate bud initiation, it apparently
permitted more buds to develop. The dependence
upon treatment during the period of vegetative
bud break in mid-May for positive re- sults is in
sharp contrast to the treatment results of Bonnet-
Masimbert and Lanares (1978) discussed earlier.
Ebell speculates that coning may result from a
sharp change in nitrogenous compounds rather
than from increased vigor after nitrogen uptake,
but he does not present data to substantiate his
hypoth- esis. Owens and Blake (1985) agreed
with Puritch when they noted that results of
fertilizer trials in stimulating initiation of
reproductive structures have been erratic, in part
because careful attention had not been applied to
the timing of the fertilizer, soils have not been
analyzed to determine possible nutrient
deficiencies, and other environmental fac- tors,
particularly soil moisture, have varied during and
after fertilizer applications. And Silen and Copes
(1972) noted that “regular applications of
fertilizer have little influence on the cyclic pattern
of good and bad cone years.” Smith et al. (1968)
presented data that show N fertilization increased
cone pro- duction of Douglas-fir by 26% in a
good year for cone crops, but none in a poor year.
In an earlier report, however, Ebell and McMullan
(1970) showed that levels of the amino acids
arginine, lysine, and ornithine, and of soluble
component of climate. According to Matthews
in trees treated with nitrate nitrogen than in (1963, p. iii), “a certain mini- mum degree of heat
those treated with ammoniacal nitrogen. In is apparently necessary for flower bud formation,
contrast, the saplings treated with probably higher than that required for the
ammoniacal nitrogen incor- porated a formation of vegetative buds.” He noted that the
higher percentage of absorbed nitrogen in importance of higher-than-average
protein than did the nitrate-treated plants.
Ebell and McMullan (1970) suggested that
specific amino acids participate in
development rather than initia- tion of
flower buds. Their hypothesis that arginine
levels may regulate flowering in Douglas-fir
is not upheld, however, at least for seedling
plants, by the data of Ching et al. (1973),
which showed that fertilization increased
free amino acids 10-fold and arginine 40-
fold. In later observations, however, no
flowers were noted on either treated or
control populations. Further studies are
needed to clarify the role of mineral
nutrients.
Steinbrenner et al. (1960) reported that applica-
tions of nitrogen (primarily) and phosphorus
stimu- lated male and female buds in a
thinned stand of 20-year-old Douglas-fir
and they reviewed papers that reported
increased cone production of several
conifers. Owens (1991, p. 256) noted that “all
things being equal, trees growing on fertile
soils produce more seed than those growing
on less fertile sites.” Masters (1982) reported
that fertilizing 12-year-old Douglas-fir seed
orchards with 224 kg nitrate per hectare as
CaNO3 resulted in 2.5 × the control number of
cones. Smith et al. (1968) found that applying
224 kg/ha of nitrate N stimulated increased
flowering on mature Douglas-fir trees. No
such effects were noted in a year with no
natural cone crop. Even though Daoudi et
al. (1994) reported that CaNO3 was more
efficient in stimulating flowers on 6-year- old
Douglas-fir cuttings than was GA4/7, the
general trend appears to suggest that GA4/7 is
more effective than CaNO 3 for flower
initiation.
Temperature
Shoots
Relatively few references discuss the effects
of temperature on flowering in conifers:
several that do, report temperature as a
134 Douglas-fir: The Genus Pseudotsuga

temperatures in June and July for flower initiation Root temperature


has been demonstrated for several species, Only one report examining the efficacy of low
however, but whether the reproductive bud root temperatures in stimulating flowering yielded
initiation occurred during those months was not nega- tive data. Zaerr and Bonnet-Masimbert
clear. Owens and Blake (1985, p. 20) reviewed (1987) found that temperatures of 5°C did not
papers that also reported on effects of higher stimulate flowering of Douglas-fir. Lavender and
temperatures during periods of flower initiation, Ching, working with two-year-old Douglas-fir
but noted that separating the effects of seedlings grafted with scions from mature trees,
temperature from those of light intensity was recorded the data found in Tables 6.1 to 6.4.1
difficult. Owens (1987) reviewed much of the Seedlings were placed in water baths in
same literature. Ross and Pharis (1987, p. 50) December 1974. In early March 1975, half of the
noted that warm thermoperiods may stimulate seedlings were maintained with an ambient soil
female flowering, but cool thermoperiods may be temperature and the second half with soil
associ- ated with increased male flowering. temperatures from 2° to 4°. All containers were
Owens (1991) reviewed several papers reporting watered at intervals. The shoot environment was
that high summer temperatures were implicated in a function of weather patterns in that the cold
increased flowers in diverse conifer species, root treatment was terminated in late October,
including Douglas-fir: “High temperatures during and all seedlings were overwintered under
floral initiation may affect metabolic processes, natural conditions; however, containers were
but we know little about these processes in insulated to prevent freeze damage to the roots.
reproductively mature trees” (p. 254). Although In early March 1976, each seedling population
data are lacking for Douglas-fir, re- searchers was halved again, with one-half of each
assume that Douglas-fir flower buds have a cold installed in the environments described above.
requirement. Nor does that suggest that the During the study, records were maintained of
chilling requirements for initiation of potential seedling phenology, incidence of strobili, and the
flow- er-bud primordia in the quiescent elongation of the shoots on the scion wood. Table
vegetative bud in early spring, nor does their 6.1 summarizes the pattern of response to the
subsequent differen- tiation during extension of treat- ment. All strobili were borne on scion tissue
the vegetative shoot, in late spring and early that developed on either the parent tree (1974) or
summer, differ from that of a vegetative bud on the non-grafted control seedlings (1975). Two
containing only vegetative primordia. One can data are of particular interest:
infer, however, that the reason for the cold
requirement of vegetative Douglas-fir buds: • The staminate strobili production was confined
to plants maintained with cold roots, and;
protec- tion against activity during an
unseasonable midwin- ter warm spell, obtains • The great majority of the ovulate strobili
equally well for reproductive buds. Studies by production in both 1975 and 1976 was
Lowry (1967), Van Vredenburch and LaBastide associated with the cold root treatment the
spring of strobilus development.
(1969) and Eis (1973a) suggest that the
meteorological sequence two years before cone Low temperatures are well known to restrict
maturity influences the size of cone crops, but the moisture movement, and therefore, such treatment
sequence does not indicate a chilling requirement may be analogous to drought. Other workers
for reproductive buds. Further, such (Ching et al. 1973, Ebell 1967, Silen 1973) have
meteorological data do not necessarily reflect reported substantial reduction of vegetative
conditions required for flower bud initiation or growth as one response to increased moisture
differentiation because abortion of buds or stress. The data in Table 6.2 indicate no such
flowers could result in a poor cone crop. Bonnet- reduction for the pres-
Masimbert (1970) and Lanares (1978) noted that
reproductive buds can develop on lammas shoots
that have had no chilling. Douglas-fir
flower buds require mild temperatures during the
spring to continue 1. DP Lavender and KK Ching, unpublished data, 1976, 1984.
development.
Chapter 6. Flowering 135

Table 6.1 Effect of seedling root temperatures on developing reproductive structures in Douglas-fir.
Control (10°C to 15°C) Cold roots (2°C to 4°C)
Number of ovulate strobili that developed on 1974 tissue of scions* 4/36 11/53
Number of staminate strobili that developed on 1975 tissue of scions 0/33 135/45
Number of ovulate strobili that developed on 1975 tissue of scions (all
but one strobilus on seedlings with cold roots in 1976) 2/33 9/45
Proportion of living grafts with strobili, 1976 1/29 21/35
Note: All staminate strobili in 1976 - percentage on plants with cold roots: 4 strobili on 36 grafts; 11 strobili on 53
grafts. Source: DP Lavender and KK Ching, unpublished data, 1976, 1984.

Table 6.2 Effect of seedling root temperatures on the growth of vegetative shoots in Douglas-fir.
Mean length of shoot (mm)
Shoot Control (10°C to 15°C) Cold roots (2°C to 4°C)
Terminal 102 97
Lateral 70 72
Source: DP Lavender and KK Ching, unpublished data, 1976, 1984.

Table 6.3 Effect of seedling root temperature upon the mineral content of scion and seedling tissue in Douglas-fir.
Mineral content
Control (10°C to 15°C) (%) Cold roots (2°C to 4°C)
Tissue N P K N P K
Seedling 0.94 0.133 0.67 1.23 0.081 0.41
Scion 1.21 0.204 1.13 0.91 0.147 0.89
Source: DP Lavender and KK Ching, unpublished data, 1976, 1984.

Table 6.4 Effect of seedling root temperature upon plant moisture stress in
Douglas-fir.
Plant moisture stress* (atms)
Date (1976) Weather Control soil temperature (°C) Waterbath seedlings (Roots @ 2°C to Control seedlings
4°C)
9 April Partly sunny, 15°C 10 13.0 12.7
6 May Warm, N.E. wind 12 13.2 11.5
10 May Cool, moist 16 11.5 6.2
19 May Partly cloudy, cool 11 13.0 6.3
* At noon measured with pressure bomb.
Source: DP Lavender and KK Ching, unpublished data, 1976, 1984.

ent populations. Further, a comparison of the 1976 The data describing the plant moisture stress of
growth of plants maintained as control both years seedlings in the spring of 1976 is shown in Table
with that of plants grown with cold roots in 1975 6.4. Surprisingly, in spite of references to
and as controls in 1976, demonstrates no “drought- induced” flowering in conifers, we
significant reduction in growth of the latter (which could find no comparable published data.
might be expected if these plants did, in fact, Although Owens (1987) suggested that apical
undergo drought stress in 1975). meristems may be more sen- sitive to small
The nutrient contents of foliage collected increases in moisture stress than previously
from both seedling and scion tissue in the fall of thought, during the summer months, Douglas-fir
1975 is summarized in Table 6.3. The erratic frequently experiences moisture ten- sions in
nature of the nitrogen content does not support excess of 25 atmospheres. So it is difficult to
a hypothesis relating this element to flowering assess the importance of the differences shown in
response. this table.
136 Douglas-fir: The Genus Pseudotsuga
The above data appear to add just one more to however, with morphology distinctly different from
the list of treatments known to induce a that of roots
“flowering response” in Douglas-fir. And, if the
mechanism of the present response is drought-
mediated, the above results are not unique. Most
of the treatments cited in this book can be
interpreted as affecting root metabolism, however.
Unpublished observations of significant
precocious strobilus production in young
plantations in the Willamette Valley, which are
sufficiently moist to permit double flushing of
many seedlings, but which support stands of
grasses known to produce materials toxic to tree
roots, pro- vide yet further evidence for this
concept. Finally, the tremendous “distress crops”
found in seed orchards when the scion and
understock are incompatible, may reflect export of
regulatory compounds from roots. If root-exported
regulatory compounds are involved in floral
initiation in Douglas-fir, however, the lack of a
flowering response in the juvenile tis- sues of the
present population indicates that such compounds
can be no more than one factor and that a strong
response to a treatment such as low root
temperatures can be expected only when this
factor is limiting.
The incidence of ovulate strobili production is
of interest because the development of such
struc- tures is believed to be determined no later
than mid-summer of the previous year (Silen
1973). If the present data are reliable, however,
the transition from vegetative bud to ovulate
bud apparently may happen until bud break.
The difference in the above results and the
nega- tive data of Zaerr and Bonnet-Masimbert
(1987) may be because these latter workers used a
minimum temperature of 5°C. Lavender and
Wareing (1972) showed that Douglas-fir seedlings
maintained at 4°C responded with good root
growth. And Lavender et al. (1973) demonstrated
that applications of GA3
to Douglas-fir seedlings maintained with cold
roots
stimulated bud break, data that suggest root-
syn- thesized plant-growth regulators may be
involved in floral initiation.
Douglas-fir seedlings can produce roots at
tem- peratures at least as low as 4°C (Lavender
and Wareing 1972). Such root growth follows
the pat- tern described by Hellmers (1963),
grown under higher temperatures. Given the both ordinary seedlings and seedlings with scion
fore- going, and the results of growth material from mature trees grafted into their
substance content of xylem exudate of Zea crowns to test the 2-year weather sequence that
mays (Atkin et al. 1973), as- saying Douglas- Lowry suggested stimulates flower production.
fir roots cultured at a range of low
temperatures for plant growth regulatory
activity might be of interest. If little or no
root activity is, in fact, a strong stimulant to
flowering, this cold root treatment may be
ideal for seed orchards in green- houses
because seedlings may be exposed to this
treatment for an entire growth season.
Although several workers have found that
cone crops correlate with weather sequences
occurring as long as 27 months before cone
maturity, but none have elucidated possible
causative effects of weather on flower
production. Ross and Pharis (1985) sum-
marized that, “studies correlating seed crops
in conifers with weather data indicate that the
proper sequence of optimal environmental
conditions (such as dry or even droughty, with
high solar insulation during the late spring or
summer before initiation and differentiation of
cone buds) for flowering may occur, but
infrequently, in nature” (p. 18).
Puritch (1972), Owens (1991), and Jackson
and Sweet (1972) reviewed reports that
indicated that warm, dry climates favor cone
production by coni- fers, and Ross (1976)
noted that the warm dry climate of Sequim,
Washington, and Saanich Peninsula,
Vancouver Island, have been correlated with
preco- cious flowering of Douglas-fir. Several
studies, Eis (1973a), Van Vredenburch and
LaBastide (1969), and Lowry (1966) have
shown correlations between weather patterns
in years immediately before heavy Douglas-
fir cone crops and the number of cones
produced. Lowry suggested that a cool July
two years before the cone crop, a moist
March-April 18 months before cone harvest,
and a warm January and mild June the year of
the crop are associated with heavy cone crops.
But Ching and Lavender (1973, unpublished
data) were unable to demonstrate that
modifications of the environment around
Douglas-fir seedlings and seedlings grafted
with tissue from ma- ture trees, (according to
Lowry’s recommendations) stimulated cone
production on the test populations. They used
Chapter 6. Flowering 137
eleva- tion over a period of 12 years at elevations
Four seedling populations were subjected to the between 2042 and 2865 m in Colorado. He noted a
prescribed sequence of weather modifications, general correspondence between levels of “staminate”
one population initiating the sequence in each of and “pistillate” flowers, a strong relation between
four successive years. Flowers appeared only eleva-
occasionally and were not related to the
treatments. In addition, Lowry (1967, p. 3) noted
that the foregoing “has produced not a definitive
explanation, but only a portion of an inferred
generalization.” He reviewed the problems of
analyzing the data set discussed in 1966. Giertych
(1987) noted a trend to move seed orchards to
warm, dry sites, and Sweet and Bollman (1972)
report that there was variation in Douglas-fir seed
production across New Zealand, but they do not
suggest a reason for these differences.
Rowe (1964) reviewed the strong effect that
fac- tors of the environment have on flower
initiation of a range of trees. Silen (1967)
demonstrated that Douglas-fir male bud
abortion increased with in- creased elevation
during the summer. Similarly, the number of
female buds fell with increased elevation,
reflecting the fact that climatic differences
associated with elevation had a negative effect
on productive buds. Enescu (1987, p. 260)
noted that, “among all the climatic factors,
humidity and temperature are the most
important ones affecting flowering and wood
production.” He cited a range of examples in
which high and low moisture and temperature
favored flowering, noting that “oak formation
of female buds is determined by low
temperatures,” and that “the transfer towards
warmer geographic regions can favor seed
maturation, early flowering and, undoubtedly,
the physical isolation against undesirable
foreign pollen . . . if the difference be- tween
the sum of annual temperatures in the natural
habitat and the place where seed orchards are to
be established is 200-300 degree days then
flower production can increase by 160%” (p.
263).
And Ross and Pharis (1985) noted that
warm,
dry sites are associated with both earlier and more
prolific flowering of Douglas-fir than the cooler,
moister sites in this species range. Roeser (1942)
presented a detailed description of the mega- and
micro-sporangiate flowers of Douglas-fir with
kill the affected part. Flowers produced
tion and delay in flowering, and the fact that immediately before death are often seen as
the flowers are most susceptible to frost “distress crops,” and they may not be a result of
damage at the time of bud break. normal stimulation and development.
Sweet and Bollman (1972) noted distinct
differ- ences in the number of seed per
Douglas-fir cones in varying collection areas
in the North and South Islands of New
Zealand, but the data do not permit conjecture
on possible climatic effects.
Greenwood and Hutchinson (1996) noted
that flowering is greater in the southern part
of a species range. They reviewed several
reports indicating that potted indoor seed
orchards frequently are profuse flowerers, but
also warned of aftereffects in seedlings grown
from indoor seed.
Eis (1972, 1973a) noted that weather over
suc- cessive summers—a cool, moist summer,
followed by a warm, relatively dry summer
the following year—was associated with cone
production in both Douglas-fir and grand fir.
These results are roughly similar to Lowry’s
(1966) and Van Vredenburch and LaBastide
(1969), discussed previously. Interestingly the
foregoing was true for both early- and late-
flush- ing trees. Sweet (1975) emphasized that
the choice of site is the most important
decision for seed orchards and that sites with
warm, dry summers favor seed production and
not necessarily vegetative growth.
Cultural treatments
The following section summarizes the effects
of a range of cultural treatments on flower
initiation.
Girdling
Numerous reports substantiate the
hypothesis that perennial woody plants may
be induced to flower by girdling individual
branches or the main stem. Presumably, this
treatment checks the basipetal movement of
substrate, making more carbohydrate
available to support development of
reproductive buds. The low soil temperature
that stimulates flow- ers on Douglas-fir may
in fact be a girdling mecha- nism that
reduces transport of carbohydrate into the
roots, leaving more carbohydrate available
to support reproductive development. Girdles,
whether applied directly or indirectly, as in
the instance of graft incompatibility, may
138 Douglas-fir: The Genus Pseudotsuga
Several trials (Ebell 1971; Skadsen 1975; time of vegetative bud break was most effective.
Wheeler et al. 1985; Bonnet-Masimbert 1982, 1987; Wheeler et al. (1985) reported that girdling Douglas-
Philipson fir definitely increased the
1990; Ross 1990; Woods 1989) all demonstrated
that girdling Douglas-fir could result in enhanced
flowering. Ebell’s trials showed that the optimum
time for treatment was about a month before bud
break. His work was particularly impressive in
that he treated one stem of trees with dual trunks
and only the treated stem responded with
increased flower production. Wood’s trials
included one poor cone-crop year when the
general promotive effect of girdling was absent.
The mechanism by which girdling affects
flowering is not known, although it is a general
phenomenon in both gymnosperms and
angiosperms. One suggestion is that the girdle
prevents translocation of photosynthate which re-
sulted in an elevated C/N ratio in the foliage. The
evidence of Ebell (1971) is inconclusive on the
role of carbohydrates in flowering and an elevated
C/N ratio contradicts the data resulting from
nitrogen fertilization, that lowers the C/N ratio.
The data resulting from Ebell’s paired stem trials
are par- ticularly interesting in that they appear to
rule out an influence of the root system unless
each stem is served by its unique portion of the
roots.
Further, according to Ebell (1971, p. 465),
“Girdling will be most promotive when applied in
years of cone crop failures, rather than years when
abundant flowering exists.” This result is the op-
posite of results reported for effects of hormones,
however. Webber and Stoehr (1998) noted that for
girdling which is more promotive than GA in in-
ducing flowering. Jackson and Sweet (1972, p.
15) reviewed several successful girdling trials and
they noted that while girdling does tend to
increase levels of carbohydrates above the
treatment, “there is no unequivocal evidence
carbohydrates play a direct role in flower
initiation.” Rather, “the important fac- tor may be
a particular biochemical situation which is
frequently associated with high carbohydrate
levels.” Sweet (1975) reviewed several flower
induc- tion treatments and noted that, though
girdling may be harmful, it can also be very
effective. Ebel (1971) girdled Douglas-fir at
different times and concluded that girdling at the
number of cones and that the cumulative beneficial. Owens and Blake (1985, p. 39) noted
effect of annual girdling was greater than that that girdling may increase cone production, but
of biennial girdling. They also found that that “results vary depending on the time of
girdling did not affect seed parameters or long application,
term health of trees. Woods (1989, p. 12)
tested various timing and types of gir- dling
and made the following recommendations:
Based on the results from this and other
studies, and experience beyond these data, the
following recommendations are made for
girdling to increase seed production:
1. Girdle from 1 to 3 weeks before
vegetative bud flush.
2. Perform a single cut into the xylem (hard
wood) with a sharp knife.
3. Proceed cautiously in local areas, with
incom- plete (90% to 98%) girdles at first,
and complete bands if experience
indicates this is necessary.
4. Girdle above at least one whorl of live
branches. This will not usually affect
cone production, and will aid tree vigor
by leaving some crown to supply
photosynthates to the roots.
5. Treat the girdle wounds with an
insecticide at the time of girdling, or wrap
the wound with several layers of
breathable cloth.
6. Fertilize girdled trees at the time of
treatment with ammonium nitrate (Ebell
and McMullan 1970) at the rate of about
300-400 kg N/ha. (Base on locale
experience.)
7. Girdle trees every second year or less.
Annual girdling may cause excessive
stress on existing cone crops and on the
trees’ ability to produce future crops.
8. Girdling should be preceded by a survey
for reproductive buds to avoid possible
damage to the current year’s seed crop.
Woods also found that effects of girdling
were found the year after treatment and that
girdling did not stimulate flowering in a
generally poor seed year.
Wheeler et al. (1985) reported that girdling
alone or with fertilizer increased flowering
over the 12 years of the study, had significant
effects on seed parameters and only a minor
effect on tree vigor. Masters (1982) reported
that girdling increased flow- ering and that
culturing trees for long internodes was
Chapter 6. Flowering 139

the method of girdling or strangulation used (the


with the vigor of the branches with the less vigor-
latter is generally not as effective as girdling) and
ous trees sustaining the greatest relative reduction
the use of adjunct treatments.” Ross and Bower
in number of cones. In contrast, the branch
(1989) showed that girdling and GA4/7 both alone
thinning reduced the cone crop.
were about equivalent in stimulating flowering on
Douglas-fir, and together they were additive. Root pruning and grafting
Bower and Ross (1985), however, showed that
Owens and Blake reviewed several reports that in-
girdling greatly stimulated staminate buds on
dicate that these treatments may increase
Douglas-fir. Bonnet-Masimbert and Zaerr (1987),
flowering in conifers and speculate that flowering
Owens and Blake (1985), and Puritch (1972)
of grafts may be caused by incompatibility,
reviewed a sub- stantial number of girdling
which, in effect, was an incomplete girdle. We
studies that vary in technique and timing; one
have seen examples of such behavior in a
conclusion of this work is that girdling should be
Douglas-fir seed orchard near Corvallis, Oregon.
timed to have a major effect at the time of
We noted earlier that root pruning or cold soil
flower bud initiation in the spring. Faulkner
temperatures could stimulate flowering. Several
(1966) described several experi- ments, including
reports have implied an influence of roots on
Douglas-fir and a range of girdles. Melcher (1960)
flower initiation, most of which suggest a
discussed effects of timing of girdling on floral
reduction in root metabolism as the causative
initiation—prior to May 30—increased the
factor. Zaerr and Bonnet- Masimbert (1987) noted
number of flowers especially female, the
that flooding the root sys- tem of potted Douglas-
following year. Girdling between end of June to
fir seedlings was associated with flowering.
July increased flowering two years later. And
Several other workers have shown that root
Sedgely and Griffen (1989) noted that “while
pruning may stimulate flowering; Masters (1982),
girdling is probably the most widely successful
for example, reported that root pruning in- creased
of the treatments designed to stimulate flower
flowering 5.2 x the control in a 12-year-old
initiation, it is no panacea. Use is limited by
Douglas-fir seed orchard, the greatest increase of
unpredictable responses relating to time, season,
any treatment. Bonnet-Masimbert and Zaerr
and cultivar, and by the long-term
(1987) reviewed trials in which reduction or
deleterious effects on tree vigor” (p. 248).
absence of root growth was correlated with
Noel (1970) presents an extensive review of
flowering in several species, including Douglas-
gir- dling, in which he discusses several trials of
fir. Interestingly, they showed that treatment with
both perennial angiosperms and conifers that have
GA4/7 blocks Douglas-fir
dem- onstrated that girdling frequently resulted in
root growth. We noted earlier that girdling was very
in- creased flowering. Although he devotes a
effective in stimulating flowering and that such
section to physiology, he does not offer a unique
treat- ment is detrimental to roots. Similarly,
reason for such results. Pharis et al. (1980) noted
Douglas-fir trees in the final states of Phelinus
that girdling branches did not stimulate flowering
weirii, a virulent root pathogen, frequently
on six-year-old Douglas-fir seedlings, and Ross et
produce heavy cone crops. Owens (1987) noted
al. (1980) reported that girdling may reduce the
that root pruning delayed devel- opment of
number of seeds per cone and, hence, may be
auxiliary apices until mid-July, when they
detrimental. And Masters (1982) reports that
developed into vegetative, reproductive or latent
girdling 12-year-old Douglas-fir trees resulted in a
buds. He speculated that midsummer environment
doubling of cone production in a seed orchard.
may have favored reproductive development.
Top pruning and branch thinning Silen (1973) reported that about 65% of sapling
Ross and Currell (1989) reviewed reports indicat- Douglas-fir trees (1.2-3.1 m tall) flowered in July
ing that top pruning both enhanced and reduced after moving on June 30-July 1 (13-14 weeks)
cone production and report the same responses for after spring flush, while only 1 of 91 control trees
Douglas-fir. They suggested the effect may vary bore cones (1972 was a poor cone year, after a
heavy crop in 1971); and that the transplanting
treatment resulted in drought symptoms and
pruned roots.
140 Douglas-fir: The Genus Pseudotsuga
Finally, Ross et al. (1985) noted that root Plant growth regulators
pruning of Douglas-fir, which resulted in a 0.3 mp Gibberellins
moisture stress at mid-day, was more effective in
These compounds were first recognized in Japan late
stimulating female flowering of Douglas-fir than
in the 19th century, when the compound found to
was applying of GA4/7. Root pruning did not
cause the “bakanae” or “foolish seedling” disease
result in increased
male flowering, however. Finally, Bonnet-
Masimbert
(1987) and Philipson (1990) found that root
pruning stimulated flowering in Douglas-fir, but
Bonnet- Masimbert and Doumas (1992) suggested
that any effects of root pruning to roots may well
be indirect.
Gravimorphism, shading
Longman et al. (1965) found that training larch
branches downwards resulted in increased flower-
ing. The same treatment has been used
successfully with horticultural species, but in
larch, which bears cones on downward oriented
branchlets, the treat- ment may have reinforced
the natural tendency. Lavender and Ching
(unpublished data, 1968) were unable to report
this effect with Douglas-fir, where bent branches
resumed a nearly horizontal orien- tation. Silen
(1973) observed that shade (13%-23% of full
sunlight) increased the numbers of female flowers
the year of application and decreased fe- male
flowering the following year. Furthermore, the
treatment had greatest effect when applied during
shoot elongation; nonetheless, the experiment left
open the earliest time of flowering enhancement.
Much of this research occurred years or
decades
ago and, though it may have been carefully done,
the design of this work generally suffers as
Romberger (1967) and Giertych (1987) noted in
their papers.
A second major effort to induce flowering uses
plant growth regulators. The compounds
apparently most successful are gibberellic acid
and analogous GA4/7. Because more than one
hundred papers de-
scribe the effects of applying gibberellins to trees,
we will consider only those that review the
literature or discuss the interaction of gibberellins
and other methods that have successfully
stimulated flowering and report dealing with
Douglas-fir.
of rice was shown to be produced by the difficulty in working with GAs and poor
fungus Fusarium moniliform Sheld., “the experimental design. In like manner, Bonnet-
asexual or imperfect stage of the ascomycete Masimbert (1982, p. 1183) reviewed research with
Gibberella fujikuroi (Saw.) Wr. T. Yabuta, GAs and made these observations:
at University of Tokyo, assigned • A positive flowering response to
the name ‘gibberellin’ to the active factor in G. specific exogenous GAs does not mean
fuji- kuroi culture filtrates in 1935, and in 1938 that they
Yabuta and
Y. Sumiki announced the isolation of two
crystaline, biologically active substances,
which they named ‘gibberellins A and B’”
(Moore 1979, p. 90). Studies of GAs in
western countries, however, did not be- gin
until after World War II. At least 86
gibberellins were identified (Pharis et al.
1992, p. 13). “However, a large proportion of
the GAs exhibit little or no biological activity,
probably because they lack the capacity to fit
a receptor molecule“ (Pallardy 2007,
p. 369). Pharis et al. (1992, p. 13) noted that
“The structural characteristics of highly
florigenic GAs depend very much upon the
plant species/family in question.” “In the
Coniferae, GAs of a wide variety of structures
are highly florigenic for Cupressaceae and
Taxodiaceae families, but only the less polar
monohydroxolated native GAs; GA 4 and GA7
can
routinely and effectively promote flowering with the
Pinaceae family. According to Pharis et al.
(1992, p. 14), GA3 has been shown to play a
role in dormancy physiology of Douglas-fir
(Lavender et al. 1973). Perhaps species that
do not require physiology that withstands
extremely low temperatures can use GA3 in
flower induction because it would not
interfere with control of cold hardiness or, as
Dunberg and Oden (1983, p. 275) noted:
“Conclusions regarding the hormonal
physiology of conifers must be based on data
from experiments with conifers. There is also
evidence that fundamental differences exist
between the Pinaceae on one hand and the
Cupressaceae and Taxodiaceae on the other. It
is therefore recom- mended that these two
groups should be treated separately when
generalizations are to be made.”
These authors also pointed out that the
conflicting results noted in reviews and
reports of gibberel- lin’s effects on flowering
before 1983 are the result of the extreme
Chapter 6. Flowering 141

naturally affect the flowering. More


convincing evidence is the endogenous Ross (1976) reported that 400 µgm GA4/7 per
increase of less polar GAs (GA4/7), as opposed branch applied at fortnightly intervals between
to stability or March and late June increased ovulate flower
decrease of GAs observed after a root pruning, production 5x over the control of 4-year-old grafts
which dramatically increased flowering of and staminate flowers, 3x. They suggested that
10-year-old Douglas-fir seedlings.” rapid conversion of GA4 (p. 185) to more polar
• Data showing relations of GA1, GA3 , GA4 , forms of GA is why relatively high amounts were
and GA9 to flowering are contradictory; such needed. The treatments were effective only on the
results demonstrate the need for metabolic clones, which flowered naturally.
studies to determine the conversion rate Wample et al. (1975) found that seedling Douglas-
between GAs, as well as quantitative analyses. fir rapidly metabolized GA to GA , GA , and un-
4 34 2
McMullan (1980) reported on an intensive known products at each of three stages of shoot
survey development—that is, bud break, shoot elonga-
of the relations of growth regulators applied to tion, and budset. This finding demonstrated that
cut twigs at the projected time of floral Douglas-fir can metabolize active forms of GA
initiation, and noted that neither bud extracts to more polar forms inactive for vegetative
from trees with a good cone production record growth,
nor those from trees
that seldom produced cones stimulated but the GA4 was active in promoting flowering in
flowering
in 10-year-old Douglas-fir seedlings. She noted Douglas-fir. Accordingly, they suggested (p. 277),
no
“therefore, Douglas-fir’s unresponsiveness (at
difference in growth regulator content of cone
least in terms of shoot elongation to exogonenous
pro- ducing or non-cone producing trees. GA 4/7
GAs may reflect either a surplus of GAs or a
stimu- lated flowering on all trees, both cone
highly ef- fective system for ‘inactivating’
producing and non-cone-producing. But the large
biologically active GAs by (a) oxidation to more
quantities used caused some tissue damage, so she
polar acidic products and b) conjugation).”
questioned whether the effect was normal and,
Ross (1976), working with equi-sized 2-year-
cited (p. 411) Reeve and Crozier (1975) to the
old Douglas-fir grafts and 4-year-old seedlings
effect that applying super physiological doses of
showed
GAs could destroy
their subcellular compartmentalization, which that non-polar GAs (GA 4/7) and, to a lesser extent,
made
difficult knowing whether a response observed GA5 and GA9, stimulated male and female flowering
was
due to the natural effect of the applied hormone on both grafts and seedlings, but that GA3 did not
or
whether it is a non-specific effect resulting from unless applied with 1AA. The trials were
abnormal chemical modification. Similarly, conducted out of doors near the droughty Sequim
Durley et al. (1975) raised the possibility that area on the Olympic Peninsula. Although none of
injecting large doses of GA20 into leaves may have the control seedlings flowered either in this
given rise to ab- normal metabolites. McMullan population did, and the author hypothesizes that
(1980) also presents the results of successive time the seedlings were therefore no longer “juvenile.”
separated samples of Douglas-fir foliage and The report of Pharis et al. (1976) is in
found no correlation with GA content and cone substantial agreement with the ones above in that
production. endogenous amounts of non-polar GAs were
Puritch et al. (1979) reported that GA4/7 high in mature,
treatment
resulted in a significant increase in cones from 8- flowering trees and was high in non-flowering,
year- GA3
old Dougals-fir seedlings and 8-to-11-year-old similarly mature Douglas-fir. They also noted that
grafts
in seed orchards over a range of sites on applying GA4/7 at the appropriate time was successful
Vancouver
Island. The seed orchard with the highest endog- in stimulating flowering on four- and six-year-old
enous production showed the greatest increase. Douglas-fir seedlings. Ross (1983) hypothesized
Pollen cones were not significantly increased by that GA4/7 in Douglas-fir is first used for vegetative
treatment. The IAA increased seed germination growth and only when this use is satisfied does
rate the chemical stimulate flowering. He further noted
and total. The GA may have changed the balance
(p. 98) that, when GA4/7 was applied early in the
of
latent vs. reproductive, vegetative buds. Pharis
and
142 Douglas-fir: The Genus Pseudotsuga
spring, it resulted in minimal flowering response Ross and Pharis (1987) reviewed many of the
and maximum vegetative growth; the reverse was papers discussed above and concluded that GAs
true if it was applied within 4.5 weeks of bud are the only growth regulators to effect elongation
break. In a series of four papers, Ross et al. in conifers.
(1985), Webber et al. (1985), and Owens et al. Ross and Bower (1991, p. 23) discussed
(1985, 1986) the authors discussed the effects of girdling and injections of GA4/7 and noted the
GA4/7 and root following:
pruning on the flowering and vegetative growth of Girdling in combination with a single stem injection
nine-year-old Douglas-fir seedlings. of the growth regulator gibberellin A 4/7 can be a
highly cost-effective treatment for enhancing seed
In the 1985 paper, Ross et al. concluded that yield in Douglas-fir seed orchards. Diminished tree
root pruning stimulated female flowers in all vigor and flowering response to biennial retreatment
families, both those with good and those with can result, however, unless tress are properly
managed to minimize the physiological stresses
poor flower- ing, but, GA was effective only on associated with treatment and the heavy cone bearing
good-flowering seedlings. Root pruning was more which follows. Alleviation of compounding water and
nutrient stresses through irrigation and possibly
effective than GA. Root pruning between 6 weeks
fertilization following treat- ment and during the off-
before bud break until early July and GA and root treatment year will speed the recovery and should
pruning were the most effective methods to enable trees to be safely retreated on a biennial basis.
(Ross and Bower, p. 23)
stimulate flowering.
Owens et al. (1985) reported that the apices They also definitely established GA and girdling
from treatment as stresses. Clearly the combination of
GA-treated and control trees were similar and fol- GA4/7 and root pruning applied at the correct time
lowed a normal growth sequence. In contrast, root is the best treatment for increasing flowering in
pruning delayed development until mid-July, Douglas-fir seed orchards currently used.
af-
ter that date, normal development of vegetative, Other treatments with GA 4/7 that have been in-
reproductive, and latent buds proceeded, with vestigated include top pruning and branch thin-
the
ning (Ross and Currell 1989). The former
greatest delay being associated with the greatest
reduced cone and pollen cone production at a
sub- sequent cone production. They agreed with
rate at least in proportion with its severity, and
Pharis and Ross (1976, p. 220) on the following:
the latter caused a response that reflected the
“Another hypothesis is that conifers utilize
change in vigor of the
endogenous GAs
preferentially for vegetative growth and it is only residual branches. Treatment with GA 4/7 increased
when environmental or other factors restrict this apices. Unfortunately, soluble carbohydrates could not be
growth that GAs are available for cone initiation.” stud- ied histochemically because they were extracted
during fixation and embedding. (Owens 1987, p. 95)
Owens (1987) largely agreed:
McMullan (1980) and Dunberg and Oden (1983) pro-
posed that cone induction treatments using exogenous
GAs enhance cone bud differentiation because GAs
are applied, and taken up in amounts far exceeding
those required for cone-bud differentiation, and
induction is not a direct morphogenic effect but a
stress effect. In the present study, GA alone had no
effect on apical size, MF, or anatomy and did not
enhance cone-bud dif- ferentiation in the trees used
for anatomical study and only slightly enhanced cones
in the general study trees (Ross et al. 1985). This
indicates the GA effect was more subtle than a stress
effect. However, until techniques are developed which
allow analysis of GAs within small apices as opposed
to whole shoots, the actual effect of GAs on apical
morphogenesis will remain unsettled. . . .
Histo-chemical tests used in this study show that
GA, RP and RP + GA treatments did not increase
total insoluble carbohydrates in potential cone-bud
production of both male and female flowers. include use of GA3, timing of treatments,
Several reviews (Pharis and Ross 1986a,b; lack of adjunct treatments.
Bonnet- Masimbert 1987; Bonnet-Masimbert • GA is more than a “stress” treatment.
and Zaerr 1987; Pharis et al. 1987; Bonnet- • Ratio of values of exogenous GA to
Masimbert and Doumas 1992, Black 1998) endogenous GA reaching bud 250-
examine the role of exogenous gibberellins in 500:1.
flowering of conifers and make the following • Cupressaceae species, which are strongly
points: stimulated to flower, utilize slowly
• Reasons for failure of GAs to metabolized GA3. Pinaceae species, which
promote flowering in Pinaceae are less efficiently stimulated, utilize rapidly
metabolized, less polar GAs.
Chapter 6. Flowering 143

• Of all treatments designed to stimulate Auxins


flowering, only root pruning in Douglas-fir is
This group of chemicals was the first plant growth
as consistent as the application of GA 4/7.
regulator studied. Indole acetic acid (1AA) was
In another trial, Bonnet-Masimbert (1982) first isolated from fungi in 1934. These substances
demon- strated that flowering of Douglas-fir with or are primarily associated with cell elongation and
without growth regulators occurred only when the were shown to be involved in the bending of
roots were inactive. While Black (1998) reviewed a stems and petioles as a result of differential
number of papers reporting either no relationship of destruction by light. 1AA has been shown to
endog- increase the flowering
enous 4/7 levels after flower inducing
response of Douglas-fir when applied with gibber-
GA treatments
or a concurrent increase in and ellin (Bonnet-Masimbert and Zaerr 1987). Ross and
GA4/7 flowering. Pharis (1987) reviewed papers which relate auxin
and ethylene to femaleness in forest trees. While
Sex expression
According to Ross and Pharis (1987), which demonstrated that GA4/7 stimulated female over
relatively little is known about the endogenous control male strobilii on Douglas-fir, but also noted that such
of sex expression (in forest trees) or its practical ma- results are not universal.
nipulation to aid pollen and crop management in seed
orchards. Well defined patterns of sexual zonation In a later paper, Ross (1990) reported that GA4/7
exist within the tree crown and shoot, and these stimulated female but not male flowering on Picea
appear to be associated with hormonal and possibly
engelmannii. In contrast, Tompsett and Fletcher (1979)
nutritional gradients. No firm conclusions, however,
are possible regarding the specific roles of different noted that gibberellin stimulated male, but not fe-
plant growth regulators in sex expression. This being male flowerings on scions of mature Picea sitchensis
as much as a problem of lack of critical study as the
maintained in a warm polyhouse. While Puritch et al.
complexity of the process itself. (Ross and Pharis
1987, p. 37)

In spite of the foregoing, they noted for Douglas-


fir (p. 39) “that strobili of both sexes differentiate
from previously undetermined auxiliary primordia
— female strobili from distal primordia that
would normally become vegetative branch buds,
and male strobili from primordia that would
otherwise abort or remain latent.” And that both
sexes apparently differentiate at the same stage of
shoot develop- ment. They reviewed considerable
evidence both for and against the theory that the
sex of a strobilius is correlated with the vigor of
the bud from which is differentiated or the shoot
where it is found. They listed several studies
Puritch et al. (1979) report 1AA did not stimulate Doumas et al. 1986; Zaerr and Bonnet-Masimbert
flowering on Douglas-fir trees. 1987; Doumas and Zaerr 1988; Imbault et al.
Cytokinins 1988; and Morris et al. 1990) reported the
following in Douglas-fir: (1) quan- tities and
Kozlowski and Pallardy (1997, pp. 311–313)
species of cytokinin varied between male, female
reported that a substance, kinetin (C 10H9N5O),
and vegetative buds; (2) levels of cytokinins
was active in promoting cell division. Since
varied with season but were highest in mid-spring;
the report from Skoog’s laboratory, cytokinin
(3) female flowers were generally associated with
has been isolated from a number of higher
low levels of cytokinin, but low levels of
plants and is apparently syn- thesized in the
cytokinin did not guarantee flowering; and (4)
roots and exported to plant shoots in xylem
high levels (5 microgram per shoot) of
sap. Meilan (1997) summarized a number of
exogenously applied iso- pentenyl adenine
reports which presented evidence that
significantly stimulated female bud formation.
cytokinins are involved in the flowering
There are no later reports expanding on the
process in angiosper- mous plants.
foregoing. Pharis and Ross (1986a,b) report ed
A number of trials at Oregon State
that cytokinins slightly enhanced GA4/7 stimulated
University, INRA (Zaerr and Lavender 1984;
flowering in Douglas-fir, but not in other species.
(1979) and Fogel et al. (1996) reported GA 4/7 stimu- Abscisic acid
lated female but not male flowering on Douglas- The compound now termed abscisic acid, or ABA
fir (Kozlowski and Pallardy 1997, pp. 312–313), was
and jack pine, respectively.
144 Douglas-fir: The Genus Pseudotsuga
originally termed “dormin” (Phillips and Wareing their populations were juvenile. Daoudi et al.
1958) and was found to be associated with dor- (1994) reports arginine levels as much as 15-fold
mancy in trees (although Zaerr and Lavender greater in trees treated with N+GA than control
1968 found it only weakly associated with trees. The treated trees flowered but not the
dormancy in Douglas-fir seedlings). Addicott control. Daoudi et al. (1994) reports that arginine
found that it was related to abscission and that proline and total free amino acids are much
when it was purified, it was identical to dormin. greater in male buds than in vegetative or female
McMullan (1980) found that ABA did not buds.
stimulate flowering in Douglas-fir, while Meilan
Polyamines
(1997) reviewed a number of papers that suggest
that ABA promotes flowering in an- giosperms by Scientists at Orléans, France reported on the rela-
antagonizing GA, but no reference is made to tionship between polyamines in buds and
conifers. flowering in Douglas-fir (Daoudi et al. 1991 and
1994; Daoudi and Bonnet-Masimbert 1998).
Carbohydrates Daoudi et al. 1994) reported that “the
Pharis and Ross (1985) reviewed a number of accumulation of polyamines in the shoots
papers dealing with flower-bud formation and accompanied bud sexualization, which suggests
noted that the environments and treatments that polyamines may constitute potential markers,
reviewed here may well have an effect on the probably rather early ones (4–6 weeks after bud
carbohydrate balance of trees. But they also noted burst), of floral initiation in Douglas-fir” (p.
that papers concerning actual carbohydrates failed 1854). They do not, however, present evidence
to establish a relation- ship between carbohydrate that polyamines are causal to flowering. Further,
content and flowering. Ebell (1971) and Ebell and conjugated polyamines are abundant in the shoots
McMullen (1970) reported that increased starch and buds of flowering plants.
levels resulting from girdling Douglas-fir were
Summary
associated with increased flower- ing. Ross and
Pharis (1987), concluded: “Thus as with floral This section has discussed many factors known to
initiation/differentiation in general, a direct influence the evocation of flowering in Douglas-
morphogenic role for carbohydrates in sex fir. Romberger and Gregory (1974) emphasize the
expression of conifers remains unproven” (p. 42). usefulness of analytical morphogenesis as a tool
to understand that developmental process leading
Arginine to flowering and suggest that without such
Ebell and McMullen (1970) related increased knowledge we will never understand fully the
levels of arginine and basic amino acids with factors control- ling flowering in trees. Giertych
flowering in Douglas-fir and suggested that (1988) argues that research to the date of his paper
accumulation of arginine may lead to cell division had really not produced efficient, reliable
favoring a continued development of sexual scenarios for production of flowers in conifers and
primordia. In a latter paper, however, McMullan agreed with Romberger and Gregory that studies
(1980) reported that applied arginine did not of development of floral anatomy were greatly
affect level of flower- ing in Douglas-fir. Pharis needed. McDaniel reviewed flowering research in
and Ross (1986a,b) cited references which report angiosperms (which is more advanced than that
negative results in flowering with applications of for gymnosperms), noting (p.
arginine to P. menziesii. Ching, KK, et al. (1973), 51) scientists should “adopt a more
reported that treatments which greatly increased developmentally oriented view of floral initiation.”
the endogenous levels of argi- nine did not cause O’Neill (1989) reviewed the range of studies
flowering in Douglas-fir, while Stewart and which tried to iden- tify “florigen” and concluded
Durzan (1965) found that increased levels of that “the multitude of biochemical changes
arginine were associated with flowering in associated with flowering are too complex to be
conifers. However, Ching et al. (1973) found no controlled by a single factor and suggests that the
such relationship in Douglas-fir possibly because riddle of flowering might best be elucidated
through a study of gene expression.”
Chapter 6. Flowering 145

The preceding comments reflect the fact that the devel-


flower evocation has rarely been the subject of opment of both male and female strobili, although
basic physiological research. One major exception the mode of action remains obscure, and (2) while
to this pattern has been the increasingly detailed there are many adjunct treatments which may fa-
study of the role of GA4/7 in the physiology of
floral evocation by Odén et al. (1995) and Pharis
et al. (1989). The former noted (p. 456) that, for
Norway spruce, “the availability of active GAs is
deliberately regulated in the specific organ
departments of the shoot, and that their
metabolism is directly influenced by various
factors including root activity, stomatal turgor
(e.g., tissue Y) and temperature,” while the latter
reported:
The flowering response to exogenous application of
GAs may imply that endogenous GAs play a part in
the flow- ering process, but does not prove it. Support
for a causal role for endogenous less-polar GAs in
cone bud differ- entiation is provided by data where
root pruning was used to promote flowering of
Douglas-fir (Pseudotsuga menziesii) over 20 fold in
the absence of exogenous hor- mone application.
Extracts of the shoots (minus needles) on which cone
bud primordia are forming showed that the root
pruning treatment had increased the concentra- tion
and/or amount of less polar GAs by 2-4 fold while
leaving the more polar GA 3 either unaffected (on a
concentration basis, or diminished on a per shoot
basis). (Pharis et al. 1989, p. 29)

They discussed further data suggesting that GA


lev- els are influenced by plant moisture content
and that environmental treatments associated with
flowering generally result in plant moisture stress.
McMullan (1980) found no relationship between
endogenous GA4/7 and flowering in Douglas-fir.
Pharis et al. (1987,
p. 72) noted that GAs of a less-polar nature appear
to play a direct morphogenic role in the promotion
of flowering in conifers, although the nature of
this role remains obscure (see also Bernier 1988,
p. 209). Additionally, Pharis et al. (1987)
concluded that “both endogenous GAs and
exogenous applied GAs of a less polar nature are
‘conserved’ in the presence of a variety of cultural
treatments which are known to promote
flowering. Such conservation would be expected
to increase their effectiveness whether flowering
was promoted by endogenous GAs, exogenously
applied GAs, or both” (p. 77).
The research summarized here suggests that (1)
for Douglas-fir (and many other conifers) the
plant growth regulator, GA4/7 is associated with
distinct cells, the larger one is the body cell, the
vor “flowering”, the most efficacious are smaller, the stalk cell, during fertilization the
those that reduce root growth. Again, the entire contents of the pollen tube are emptied into
mechanisms for this possible reaction are the archigonium. He concluded, “The account
unknown. It would be interest- ing to repeat here given of the gametophytes of Pseudotsuga
the low temperature exposure of roots with makes
the addition of examination of apices (Owens
1987) together with analyses of GA content.
The juvenile-mature phase change remains a
“black box,” although Mellerowicz et al. 1995
(p. 443) suggest that there is circumstantial
evidence of changes in chromatin
organization or both, during maturation.

Embryogeny
Embryogeny and pollen are the subjects of a
vast botanical literature, particularly for
angiosperms. So, my discussion will be limited
generally to Douglas- fir. Much of the
material published before 1972 is reviewed in
greater detail in Allen and Owens (1972).
Lawson’s paper (1909) is the earliest
thorough discussion of the embryogeny of
Douglas-fir. He describes the male and
female gametophytes in detail, noting that
the pollen is globular in form with a distinct
well-developed exine, but no wings to
provide the buoyancy. Within the
microspore there are two prothalial cells, two
large free nuclei, one centrally located and
termed the generative nucleus, the second,
the tub nucleus. The research was started too
late (30 March) to permit early de- scription
of the megaspore, which contained large
quantities of sap and little cytoplasm, and
three or four nucleii. Just prior to fertilization
a ventral canal cell and an egg form. The
nucleus of the egg becomes greatly enlarged
and descends to the center of the
archegonium. The pollen grains lodge in the
upper chamber of the micropyle, where they
are caught on the extended surface of the
micropyle prior to the folding of this
structure, which is typical of Douglas-fir.
The pollen grains do not move to the nucellus,
but germinate in place, extending a long
pollen tube (another characteristic of
Douglas-fir) through the nucellus to the
archegonium. Just prior to the extension of the
pollen tube, the generative nucleus divides,
which results in the organization of two
146 Douglas-fir: The Genus Pseudotsuga
it clear that this genus is not closely related to the The megaspore membrane makes its appearance at
a very early period, and although quite thin at first it
genus Tsuga. And considering the state of devel- increases in thickness with the growth of the prothal-
opment of the various vestigal and semi-vestigal lium, and eventually becomes very conspicuous. In
structures present, the view that the Abietineae are the mature stages it surrounds the prothallium
except in the region of the archegoria. In this region is
the most ancient group of the Coniferales is very it entirely absent, in this regard differs quite markedly
much strengthened” (Lawson 1909, pp. 177-178). from Tsuga. With the increase in the size of the
central vacuole, and the consequent formation of the
Lawson’s (1909) work was limited by the
parietal layer of cytoplasm, free nuclear division
instru- ments available at that time and by the fact continues for some time. The parietal layer now
that his observations of the microspore increases in thickness, and the primary prothallial
cells are formed in the ordinary way. These latter
commenced after the initiation of development of structures elongate in an inward direc- tion, and
these structures in early spring. His observations gradually close the central vacuole. Free nuclear
and conclusions are reproduced here. division now takes place within the primary
prothallial cells, before cross-walls are formed to
The microspore at the time of pollination is globular organize perma-
in form and differs in appearance from that of the nent prothallial tissue.
majority of other Abietineae in the entire absence of The archegonia originate as superficial cells at the
bladder-like appendages. apex of the prothallium.
The mature microspore contains four cells. Two of They are generally four in number, and each is
these are represented by the fragmented remains of en- veloped by a single layer of nourishing jacket-
two vestigial prothallial cells, and the other two cells. There are generally two tiers, but frequently a
represent the tube and generative cells respectively. single tier of neck-cells.
Owing to the peculiar form of the micropyle, The archegonia are separated from one another—
which has a stigmatic surface at the mouth, the pollen especially in the region of the necks—by several
grains fail to reach the apex of the nucellus, but are layers of sterile prothallial cells, and each is provided
caught at the mouth of the micropyle and here with a separate archegonial chamber.
germinate. A distinct ventral canal-cell is formed as a result
This pollen-receiving device and the formation of of the division of the central cell.
pollen-tubes so far removed from the nucellus is The membrane of the ventral canal-cell persists up
unlike anything yet reported for the Abietineae, and is to the time of fertilization.
evidently a novelty as far as the Gymnosperms are The fusion of the sex nuclei takes place in the
concerned. middle of the archegonium.
With the first appearance of the pollen-tube the The female is many times the size of the male.
generative nucleus divides, and as a result of this divi- The first segmentation-spindle is formed within
sion two distinct cells are organized, one of which is the area bounded by the membrane of the fusion-
considerably larger than the other. These are the nucleus. It is, however, of cytoplasmic origin. One or
body- and stalk-cells respectively. more dense masses of cytoplasm are carried into the
The pollen-tubes grow down the micropylar canal egg-nucleus by the sperm-nucleus.
and attain a considerable length before the nucellus is The first division is very soon followed by a
reached. second, and the four free nuclei thus formed pass to
The tissue of the apex of the nucellus disintegrates the base of the archegonium.
in advance of the approaching pollen-tubes, so that After the division that follows, cell-walls are
the latter structures find little or no obstruction in their formed separating the nuclei.
path towards the archegonial chambers. Eventually the pro-embryo consists of three tiers
The division of the body-cell results in the of cells and one tier of free nuclei. The lowermost of
formation of two male nuclei of unequal size. these becomes the embryo proper. The middle one
The entire nuclear contents of a pollen-tube are becomes the suspensor, and the next one the rosette.
dis- charged into one archegonium. (Lawson 1909, pp. 176-177)
There are probably three megaspores resulting
from a single mother-cell. Two of these are abortive Buchholz (1920, 1926) discussed the question
and one functional. of simple polyembryony versus cleavage
Upon the enlargement of the functional megaspore
free nuclear division takes place, and this is followed
polyembry- ony in conifers. He presented
by the formation of a large central vacuole. evidence that simple polyembryony, wherein a
Completely enveloping the growing megaspore single embryo develops in each archegonium,
there is a single layer of large sporogenous-like cells
which are closely packed together. This layer of cells, indicates that this type of embryo development is
although single at first, soon becomes several layers evolutionarily advanced over cleavage
thick, and eventually becomes quite loose and sponge- polyembryony, wherein more than one embryo is
like—with numerous inter-cellular spaces—as the young
prothallium increases in size. This tissue is regarded present initially in each archegonium. He further
as sporogenous in origin and tapetal in function. argued that simple polyembryony is as- sociated
with siphonogamy, or the development of
Chapter 6. Flowering 147
nation of Douglas-fir is described in detail as follows:
pollen tubes and the lack of functional “rosette”
cells. On this basis, Douglas-fir is the most
advanced of the Abietineae, Pinus, the most
primitive. He noted that, while Douglas-fir has
simple polyembryony, there can be as many as
eight archegonia in each ovule, and hence there is
the possibility for signifi- cant competition at the
embryo stage, and that the successful embryo is a
product of superior egg and pollen tubes. It is
interesting that, in his discussion of the possibility
of pollen tubes and motile sperma- tozoa some
four decades later, Christiansen made no mention
of Buchholz’s work. Gravatt et al. (1940) noted
that one lot of Douglas-fir seeds had 5 of 1,214
multiple seedlings, a second lot, 47 of 1,174.
In a series of papers, Professor J. Doyle
(Doyle 1926, 1945; Doyle and O’Leary 1935)
discussed the ovule of Douglas-fir and its
pollination in detail. He agreed with Lawson
(1909) that the top of the micropilar tube is a
stigmatic surface, which when it invaginates,
brings the pollen near to the mi- cropilar tube
on the numerous hairs on its surface (Doyle
1926). However, he disagreed with Lawson, who
thought that this method of pollen capture was
unique to Douglas-fir, and he noted that Larix
leptolepsis has a similar structure—the
difference between the species being that
Douglas-fir pollen germinates while attached to
the hairs, whereas in larch, the pollen
germinates after falling from the stigmatic
surface to the nucellus, where it develops. He
described how, after pollen has landed on the
stigmatic surface, this structure invaginates,
holding the pollen grains in a depression near
the nucellus. By so doing, the stigmatic surface
effectively closes the micropilar canal. Doyle
(1926) observed the following:
The micropilar differentiation is more elaborate than
in Larix. Just above the nucellus the micropilar canal
con- tracts to a very narrow slit. Above this, it is
continued as two lips. On the side of the ovule, toward
the centre of the scale, these lips are joined at their
base, like the fused petals, of say, a Veronica; on the
side towards the scale edge, they are completely
separate. Both lips are differentiated, but the outer one
much less so than the in- ner, which, slightly wider
than the outer, is much larger, completely overtopping
it. The two lips, at pollination are closely pressed
together and project from the narrow ovular top as
white turgid mass. (Doyle (1926, p. 177)

In a later paper (Doyle and O’Leary 1935), the polli-


between the ovule and the scale, germinating as they
In Douglas-fir, it is well known that the pollen lie; and others which had been lodged so far back on
remains at the top of the micropyle, more or less the stigmatic expansion that they were not enclosed
entangled by the hair-like outgrowths to be seen by its subsequent incurling, may frequently be found
on the inside of the inturned rim and which have in early growth stages. These grains also seem to
resulted from the col- lapse of the stigmatic show no tube growth until cone inversion is complete;
expansion. Some of the grains, however, are to but they
be found later a short distance down the
micropyle, but none get at all near the nucellus.
Lawson (1908) seems to suggest that they are
prevented from reaching the nucellus by the
contraction and kink, which appear some
distance down the micropyle neck, and which
reduce the canal to a slit-like passage. This,
however, can hardly be the case. In the first
place, the narrowing of the micropyle is not as
marked as suggested by Lawson, a space quite
wide enough for the passage of several pollen-
grains being left. In the second place, the ovules
in many of the cones are actually inverted
spatially at pollination, so that only during the
subse- quent inversion of the cone could the
grains fall down the micropyle. Even if the cones
all lay horizontally the grains in some of the
ovules could only fall a short distance down the
micropyle as a result of the oblique lie of the
upper part of it; and the subsequent complete
inversion of the whole cone, which brings the
actual ovules into an erect position, is so slow,
and the grains so small, that, moist as they are, it
is clearly unlikely that they can slip any distance
during the process. During the closure, however,
some doe become detached from the inturning
rim and slip a little along the micropylar canal.
The inverted position of so many ovules
combined with the angle, but not the narrowness
of the micropylar canal, puts a limit to their
further passage. The grains, as is also well-
known, ultimately germinate in situ sending their
pollen-tubes through the lumen of the micropyle
fairly directly to the nucellus. As the pollen-
grains of the Douglas-fir are among the largest in
the Coniferae, tubes, dissected out at this stage,
form excellent material for the examination and
demonstration of early growth of the Abietinean
pollen-tube, the body-cell being very massive
and distinct.
There is, however, a considerable delay in the
germi- nation of the grain. The young cones,
erect or sub-erect during pollen reception, remain
so for perhaps a week, and then slowly begin to
bend downwards, becoming markedly inverted,
the process being completed in about three weeks
from the original pollination. As far as can be
ascertained no tube growth begins until a few
days after inversion is complete. By this time, the
whole micropyle tube has increased
considerably, especially in thickness, the kink
referred to has become largely obliterated, and,
except in the case of grains tucked very
completely under the remnants of the inturned
rim, the tubes can grow directly downwards, as
the ovules now, owing to cone inversion, stand
spatially erect. It is tempting, of course, to
suggest that some gravity stimulus is here
effective, but as no experimental work has yet
been carried out any suggestion is out of place. It
is common also to find grains which have lodged
on the scale, usually those resting in the groove
148 Douglas-fir: The Genus Pseudotsuga
never develop for any length of time and die in microsporangium consists of one layer of cell”
about seven to ten days. During the growth of the
functioning tubes within the micropylar canal the (Gifford and Foster 1988, p. 423).
cells of the tissue are very succulent and turgid, and
there is possibly an exudation of fluid within the
micropyle, but the whole question of the growth of
the grains, both in the ovule and in vitro, has still to
be investigated. (Doyle and O’Leary 1935, pp. 199-
200)

In a later paper, Doyle (1945, p. 47) observed


that “an inverted ovule, an exudation of fluid from
the micropyle, and a two winged (pollen) grain or
some variant of it, are taken to be basal in the
Pinaceae family.” As he also noted, however,
there are at least four variations occurring in the
family of this basic pollination pattern:
The Larix—Pseudotsuga type. This type shows, then,
an extreme advance on the Picea orientalis type. The
stig- matic function of the micropylar area has here
become dominant, a large swollen stigma developing,
which covers the opening. There is no fluid secretion
at the time of pollination, and it is clearly of interest
in relation to this that the large pollen grains caught
externally on the stigma have completely lost their
air-sacs, and with them the capacity to float. These
grains have become so specialized from the primitive
type that they show no real vestige of the germinal
furrow the wall being uniform over practically the
entire surface. Inversion and invagination of the
stigma brings the grains within the ovular cavity.
Pseudotsuga probably shows a slightly more
advanced stage than Larix. The cones show no
geotropic reaction at pollination but stand at any angle
between erect and horizontal depending on the
directional lie of the actual bud and branch that bears
it. The characteristic drooping position of the older
cones is one which is gradually as- sumed long after
pollination. (Doyle 1945, p. 49)

Many of Doyle’s observations have been


confirmed
by those of Allen (1943; 1946a,b; 1947a,b).

Microsporangiate strobilus
Gifford and Foster (1988) observed the following:
The microsporangiate strobili of most conifers are
rela- tively small, commonly measuring only a few
centime- ters or less in length. [I]n all coniferous
species, the
microsporangia develop on the lower surface of the
sporophylls … in the Pinaceae the number is
constantly two. The initial cells of the microsporangia
of conifers are asserted to lie below the surface or
epidermal layer of the microsporophyll. (Gifford and
Foster 1988, p. 422)

They noted, however, that in two membranes of


the Pinaceae, the sporangial initials are
superficial, as reported by Allen and Owens
(1972), and that, “at maturity, the wall of the
Singh (1978), who described the when shed, the pollen “is no doubt quite variable
development of the gymnosperm in stage of development,
microsporangium in detail, concurred with
Allen (1946b) that the sporogenous cells are
differentiated deep within the two spo-
rangia (Singh 1978, p. 9). General
observations on microspores (pollen grains)
were made by Blackmore and Knox (1990):
Sporogenesis, whether leading to the formation
of iso- spores, microspores or megaspores,
begins with the meiotic division of sporocytes
(spore mother cells) in specialized organs termed
sporangia and culminates with the germination of
mature spores. The ontogenetic programme of
the haploid generation following meiosis varies
enormously between the major groups of land
plants. (Blackmore and Knox 1990, p. 2)

Pennell (1988) noted that a major difference


in the de- velopment of microspores (pollen in
gymnosperms and angiosperms) is that in the
former meiosis lasts much longer (almost 2
months in Pseudotsuga), whereas in
angiosperms two or three days is more
common. Further, almost all the extended
period in conifers extends in the prophase.
Pollen grains, the multicellular microspores of
seed plants are equivalent to highly reduced
gametophytes con- sisting ultimately of a
single vegetative cell or two sperm cells.
Owens and Molder (1971) gave a detailed
discus- sion of the development of Douglas-
fir pollen. They noted (p. 1260) that meiosis
begins in October and proceeds through to the
early prophase stages of the pachytone. It is
arrested until late February when meiosis is
generally completed and the microspore
region has several hundreds of haploid, thin
walled microspores each containing a haploid
nucleous and starch. The microspores then
develop, until early in April, each pollen grain
consists of a wall enclosing five cells, two
small lens-shaped prothallial cells, a stalk cell
and a large tube cell.
In their detailed discussion of the
development of the grain, Owens and Molder
(1971) found that, contrary to Christensen
(1969a,b), there is definitely a generative cell,
as in other conifers; they argued that the
normal sequence of pollen grain development
in Douglas-fir is the same as in Pinus. They
found that at pollination, about 50% of the
grains are at the five-cell stage, mature pollen
grains normally have five cells, and that,
Chapter 6. Flowering 149
no wings or bladders. When wings are present, there are
being somewhere between the three and five usually two, as in Pinus, but pollen grains of several
genera in
celled stage” (pp. 1263, 1265). The paper
confirms previ- ous suggestions by Lawson
(1909) and Allen (1943) regarding the
development of Douglas-fir pollen grains to the
five-cell stage. The structure of the pollen grain
wall or sporoderm is similar to that of other
conifers (Chamberlain 1935).
Allen and Owens (1972) found that when
pollen grains are shed, they consist of five haploid
cells and are actually an immature male
gametophyte. Two processes are involved in
pollen production: microsporogenesis “and the
development of the mature multicellular pollen
grain or male gameto- phyte from the one-celled
microspore” (Allen and Owens 1972, p. 45). They
described Douglas-fir pollen production in detail:
Meiosis is usually completed in Douglas-fir by the
end of February and each microsporangium is filled
with several hundred tetrads of haploid microspores
borne in a watery fluid. Each single-celled, haploid
microspore develops into a pollen grain during
March, within a few weeks following meiosis. Each
microspore of the tetrad is angular and the four fit
compactly together to form a sphere within the
microspore mother cell wall. Each microspore
contains a single, haploid nucleus and little starch. For
the first three weeks following meiosis no cell
divisions occur within the microspores. The cell wall
of the microspore rapidly thickens equally on all
surfaces. Microspores enlarge slightly but remain
together within the microspore mother cell wall.
Rapid accumulation of starch occurs until the
cytoplasm of each microspore is densely packed with
large starch grains. . . .
The mature pollen grain of Douglas-fir at the time
of pollination usually consists of five cells: two small
lens-shaped prothallial cells; a stalk cell; a body cell,
and a large tube cell. Mature pollen is 90-100 m in
diameter, approximately spheroid, and is usually
indented on one side. Unlike some other conifers
(Chamberlain 1957), it lacks bladders (wings) and
conspicuous pores or furrows (Barner and
Christiansen 1962). The thin microspore cell wall
thickens during microspore enlargement and pollen
grain development. The exine or outer wall layer is
thick and its surface is very smooth except for a very
faint triradiate ridge indicative of the mutual contact
among members of the spore tetrad. The intine, or
inner wall layer, is about equal in thickness
(approximately 2 m) to the exine. (Allen and Owens
1972, pp. 54-55)

In their conclusion, Allen and Owens (1972)


compared
Douglas-fir pollen to that of other gymnosperms:
Pollen grain development and structure are variable
within the conifers (Chamberlain 1957; Bierhorst
1971). All conifers are wind pollinated and two-thirds
of the genera, including Pseudotsuga, have pollen with
tetrad. During the tetrad period, a microspore
the Podocarpaceae have two to six wings. The
wall of the pollen grain has two distinct layers –
surface coat is deposited and the two exine layers
the exine and intine. Usually the exine is thicker, (the outer exine and inner intine) are initiated”
but in Douglas-fir they are equal. Like Pinus, (Kurman 1990, p. 157). Two decades after the
Douglas-fir is an example of the prevalent course
of pollen grain development in conifers.
work of Owens and Molder (1971) for Douglas-
Prothallial cells are a constant feature of the fir, Blackmore and
Pinaceae, while in all the Taxaceae, most of the
Taxodiaceae and many of the Cupressaceae, they
are lacking. Wherever there are no prothallial
cells, pollen development oc- curs as in the
angiosperms and is interpreted as being more
advanced. In most of the Podocarpaceae and all
of the Araucariaceae, many prothallial cells
develop. This results from division of two or
three prothallial cells formed from the
microspore nucleus. The male gametophyte of
different genera can be found at various stages of
development when the pollen is shed but, for a
given species, this will vary little, if at all.
Douglas-fir pollen may be shed at the four-celled
stage but is more commonly shed at the five-
celled stage. In some species of Cupressus and
Juniperus, the uninucleate microspore is shed
and subsequent development occurs within the
seed cone before pollen-tube formation. . . .
The pollen cone enlarges during the month,
usually March, of pollen-grain development.
Elongation of
the pollen cone within its bud scales begins at
the end of February, at about the same time as
pollen develop- ment. Bud scales do not
enlarge but the pollen cone does, forcing the
bud scales apart. Pollen-cone growth results
from elongation of the entire cone axis, which
causes separation of the microsporophylls and
consider- able elongation of the stalk at the
base of the cone. No apical growth occurs and
no new microsporophylls or microsporangia
are initiated during this period of pollen-cone
growth. Bud burst, resulting in shedding of the
pollen, generally occurs early in April. (Allen
and Owens 1972, p. 58)
According to Adams (1982), while there
are differences between clones of Douglas-fir
in pol- len production, these differences are
not great. He found that the range of weight in
pollen grains was from 2323 to 3112 grains
per milligram and that the number of
microsporophylls varied from 52 to 89 per
pollen cone with a mean of 74. Kurman
(1990) noted that the great majority of the
papers describ- ing the ultrastructure of pollen
development are concerned with
angiospermous pollen and, in the reports
which do report coniferous pollen, there is a
wide range in the timing of development of
pollen grains with species. She observed that
“the earliest stages of pollen wall
development occur as soon as the micro
spores were separated from one another in the
150 Douglas-fir: The Genus Pseudotsuga
Knox (1990) summarized these processes for for years been in common usage. The use of male and
female when referring to sporophytic (spore-
plants in general as follows: producing) structures such as cones is incorrect. It is
Not only are the processes of microsporogenesis only the gameto- phytic (gamete-producing)
gener- ally continuous, but a number of processes structures, pollen grains and female gametophytes,
usually occur concurrently (Blackmore et al. 1988). that can correctly have a particular sex attributed to
Thus as microspore wall deposition proceeds, them. They are the only structures that ultimately
programs of cytoplasmic and nuclear activity take produce male and female gametes. The misuse of
place within the microspores and sporangial male and female in this manner, however, seems too
development progresses around them. This complex well-established to be easily overcome.
situation makes it impossible to recognize a single Pollen cones vary in shape from globose in many
series of discrete developmental stages that will serve Cupressaceae to the more familiar cylindrical shape in
as the basis for comparison between all plants. This most other conifers. Their appearance is largely deter-
problem is not confined to the study of microspo- mined by the nature of the microsporophylls. Some
rogenesis but is confronted by every systematic and ap- pear very leaf-like, as in certain species of
evolutionary comparison of ontogeny. A series of Araucaria and Picea, while in most other conifers,
major developmental landmarks can be established for including Douglas- fir, they are very reduced, blunt,
the discussion of microsporogenesis but it must be sac-like structures. In Douglas-fir, the leafy blade of
recog- nized that differences in timing are less the microsporophyll is much reduced and only the tip
significant than differences in the developmental is apparent beyond the swollen microsporangia.
processes involved. Blackmore and Knox (1990, p. 3) The microsporangia are
commonly borne on the abaxial (lower) surface of the
In their lengthy monograph on pollen, Stanley sporophyll. The dominant number of microsporangia
and Linskens (1974) detailed pollen cell-wall is two, being considered throughout the Pinaceae, but
many other coniferous species have more. (Allen
forma- tion and composition (primarily for and Owens 1972, p. 39)
angiosperms), finding that accumulated
temperatures are an im- portant determinate for They described the phenology of the Douglas-fir
time of pollen dehiscence (in gymnosperms), pollen cone as follows:
following the simple parting of the microstrobili Unlike seed cones, the life cycle of the pollen cone is
only slightly more than one year in duration. During
sporephylls, and that, generally, pollen size is this time they become conspicuous for a few weeks
related to chromosome number and temperature. during pol- lination in the spring. Pollen-cone buds
are initiated by the first of April as undetermined
They noted that Douglas-fir pollen is among the axillary bud primordia
largest of conifers (p. 27). Further, they reported — the same as vegetative and seed-cone buds.
that 27°C temperature accelerated dehiscence of Generally, by early June, pollen-cone apices can be
distinguished from other apices only by histochemical
Douglas-fir pollen by 4 weeks over natural means, but by early July, a distinction can be made by
conditions (p. 54). In reviewing the role of growth carefully removing the bud scales and observing the
regulators in pollen physiology, they found that apex. At
that time microsporophylls begin to be initiated and
growth substances diffusing from pollen may the apex continues to enlarge. Microsporangia
stimulate maturation or receptivity of the egg cell” develop
and that “growth substances can control tube on the microsporophylls throughout the summer and
all microsporophylls and microsporangia are formed
exten- sion in many ways, one of which is by early fall. Pollen-cone buds can usually be distin-
facilitating wall growth” (p. 258) guished externally from seed-cone buds by the end of
July. Buds appear dormant by early December.
The pollen cone Although the earliest stages of meiosis begin early in
the fall, ma- ture pollen does not form until spring.
In their excellent review, Allen and Owens (1972) Cones enlarge during March and burst through the
noted the following, with respect to the microspo- bud scales about the first of April. Pollination may
continue on a tree
rangiate strobili of Douglas-fir: for about 2 weeks. The pollen cones become
All conifers bear pollen cones in the form of simple completely dry and usually fall from the tree within a
strobili. The pollen cone, a simple strobilus, consists few weeks. This cycle is similar to that in most other
of a single axis bearing a series of usually spirally conifers in that pollen cones are initiated
arranged pollen-forming appendages, the approximately 1 year before pollination. The precise
microsporophylls. No structures form in the axils of time of pollen-cone initiation and pollination,
the microsporophylls so it is not a compound however, may vary considerably among species.
structure. The pollen cone has (Allen and Owens 1972, pp. 40)
been referred to as microsporangiate strobilus, stami-
nate strobilus, male strobilus, male flower or male Pollen-cone development is detailed as follows:
cone. Microsporangiate strobilus is morphologically Early development of the pollen-cone bud involves
the most accurate but seldom-used term, and male fre- quent cell divisions in all planes, which causes a
flower has
small, dome-like apex to become visible in the
leaf axil. Bud
scales begin to be initiated within a few days
after lateral bud initiation, when the bud
primordia are only a few
Chapter 6. Flowering 151
upturned tip that contains a large air space. The
cells in height......The apex enlarges more slowly than mature microsporophyll then consists of a reduced foliar
potential vegetative and seed-cone apices. As a result, structure with a single vascular bundle anatomi- cally similar to
the zonation pattern described for pollen-cone apices . that in leaves and bracts. On the abaxial
..
does not become apparent as early as in the other
types of apices. The apex gradually assumes a more
conical ap- pearance, while bud scales continue to be
initiated along the flanks of the apex, elongate,
overarch and enclose the apex. Toward the end of
the period of bud-scale
initiation, about mid-July, the pollen-cone apex shows
a zonation pattern similar to but less distinct than that
described for vegetative apices Bud scales enclosing
the pollen cone are fewer in number than in vegetative
buds. Whether this has any influence over pollen-cone
development or is simply another manifestation of de-
velopment has not been determined.
The base of the developing bud, where it attaches
to the branch from which it originated, broadens
during apical enlargement. This results from both the
broaden- ing of the pith and the formation of a
meristematic region, a receptacular meristem, in the
cortex of the developing bud. Similar growth occurs
in all lateral buds, but to a lesser extent in pollen-cone
buds. as a result, pollen-cone buds attach less firmly
to the branch. It is this region of the cone axis that
breaks so readily after pollination and this partially
explains why Douglas-fir pollen cones seldom remain
long on the tree following pollination.
Microsporophyll initiation begins after all bud
scales have been initiated, about mid-July, and is
complete by early fall. The pollen-cone apex at the
onset of microspo- rophyll initiation is slightly
smaller, about 200 µ high and 200 µ wide, than seed-
cone or vegetative apices at the same stage of
development. The entire bud
enlarges during microsporophyll initiation but
because microsporophylls are initiated in rapid
succession up the flanks of the apex, the apical dome
is continually “used up” and gradually diminishes in
size. When
microsporophyll initiation is complete, the apex is re-
duced to a flattened dome — 60 µ high and 200 µ
wide. The method of initiation and early stages of
develop- ment are very similar in microsporophylls
and leaves. . .
. Microsporophylls are first evident as a group of
surface cells that elongate radially, divide periclinally
and pro- duce a symmetrical, hemispheric
protuberance — the primordium. Continued divisions
become organized in a predictable manner and, as in
leaves and bracts, an apical form of growth occurs.
Microsporophylls, however, do not elongate nor grow
upward as much as leaves or bracts but form blunt
foliar appendages. This is a result of equal growth on
both surfaces and especially abaxial cells dividing in
all planes, resulting in increased volume rather than
causing elongation. The microsporophyll thus remains
short, more massive, and stands more perpendicular to
the cone axis than bract or leaf primorida. Due to
continued rapid growth, the
abaxial side swells and gives rise to two
microsporangia, one on either side of the midrib,
completely joined along their inner surfaces. The
midrib extends beyond the
adjacent microsporangial regions and forms a pointed,
(1969) de- scribed the development of the pollen
surface are two large, medianly fused
micrsporangia that give the structure a
grain in detail, confirming his earlier work to the
blunt, pouched appearance. . . . effect that meiosis takes place in March (in
Microsporangial initiation occurs Denmark), meiosis during the following 2–3
when the new- ly formed microsporphyll
is only about 75 µ long. Microsporangia
weeks, and pollination in May. The pollen grains
originate by the division of several su- are globular at pollination time and elongate
perficial (protodermal) cells on the during the next 3-4 weeks. During this period, the
abaxial surface of the microsporophyll.
(Allen and Owens 1972, pp. 40-41)
embryonal cell moves toward the middle of the
pollen grain. At germination time (mid-June), the
Christiansen discussed the body cell dissolves and two male cells appear.
development of Douglas-fir pollen and Christiansen (1969) noted, however, that the
the fertilization mechanism in a series of struc-
reports (Barner and Christiansen 1962;
Christiansen 1969, 1972). He made the
following observations in the first
report:
(1) When the pollen is placed in water, it
throws off the exine and elongates until it
is 550 m. However, this response appears
to be a purely mechanical process as the
grain does not release male gametes at
this time that and may not be viable.
Barner and Christensen (1962) note that
they have never been able to achieve true
germination in vitro and speculate that the
pollen requires some special stimulus
produced by the nucel- lus when the egg
cells are receptive. (2) They agree with
Lawson’s observations of the
development pollen grains but note that
their material limited their ability to
thoroughly describe the ontology of
pollen. (3) Three to four weeks after
pollination, the grains swell, cast off the
exine, and elongate. Seven or eight weeks
after pol- lination, pollen grains are
transferred to the nucellus top from the
stigmatic flap (they do not know how this
is achieved, but speculate that a
pollination drop exuded from the nucellus
may be the mechanism – later work by
Allen and Owens 1972, refutes this
hypothesis). (4) Non-motile gametes are
discharged through a pollen tube into the
nucellus. (5) The beginning of the
receptive period of the megagametophyte
is characterized by the rupture of the bud
scales covering the inflorescence. (6) All
stigmatic flaps on a megagametophyte do
not open simultaneously, but 5-6 days
after the receptive period begins, most are
open. (7) The basal portion of the bracts
is largely rectangular and the ovuliferous
scales are about seven times their size at
the beginning of the receptive period/at
the end of this period, about 12–14 days
after the opening of the megagametophyte
bud. The end of the receptive period is
characterized by the collapse of the
stigmatic flaps. (Christiansen 1962)

In a subsequent report, Christiansen


152 Douglas-fir: The Genus Pseudotsuga
ture of the cell and the mitotic divisions of the 61 to 96 with an average of 77, demonstrating
pollen consider- able variation. (3) While microsporangia
commonly bear two microsporophylls, some were
grain are extremely difficult to analyze: found with one. (4) Average number of pollen grains
The spermatozoid of P. menziesii evidently is a per microsporangium was 462 (258–724). The
unicel- lular, multiciliated organism with a powerful number was greatest at the base of the microsrobilis.
locomotor apparatus, a neuromotor system and organs Each microstrobilus produces ap- proximately 59,600
for orienta- tion, etc. (Christiansen 1969, p. 98). pollen grains. There are about 3 mil- lion pollen
grains per gram. (Ho and Owens 1974, p. 561)
The elongated part of the pollen grain is sometimes
termed “pollen tube”, but in view of the fact, that it Dr. George Allen authored a series of reports
does not grow into a style or apex of nucellus, but (Allen 1946, 1947) discussing in great detail the
sprouts a special tube, through which male cells are
discharged, it seems doubtful if this term is justified. anatomy and ontology of the mega- and micro-
It may also be questionable whether the special short sporangium and embryology of Douglas-fir and
tube is a pollen tube in the usual sense of the word; it their relationship to published data for other spe-
is not clear if it is always a grown tube, or a tube
made by chemical means. (Christiansen 1969, p. 101). cies. The detail is, perhaps, not appropriate for
this volume, so we will limit the reference to
It is suggested that at germination the spermatozoids,
remaining inside the membrane surrounding the body
quoting his summaries and noting his major
cell complex, are propelling themselves and the points. After a lengthy discussion of the origin
contents of the membrane through a short (pollen) (upon the shoot) of the microsporangium, he
tube into the apex of the nucellus and on to the
vicinity of the egg cell. (Christiansen 1969, p. 103). concluded,
The microsporangium of Pseudotsuga is a superficial
In the final report, Christiansen (1972) continued structure and the generally accepted concept of a
his description of the pollen grain and pollination hypo- dermal origin for the sporangium in conifers is
open to question. The sporangium cannot be traced to
mech- anism, primarily of Larix, but also of a single cell or group of cells. It has an ontogeny
Pseudotsuga. He argued that, as far as pollination similar to that of the lower tracheophytes, and, unlike
and pollen grains are concerned, the two species the angiosperm microsporangium, is at no time
invested by a protoderm or epidermis genetically
are closer to the Ginkgoales than to the continuous with that of the shoot. Bower’s (1896)
Coniferales. Christiansen confirmed earlier work concept that a relationship exists between
to the effect that Pseudotsuga does not produce a stratification of the shoot apex and stratification of the
young sporangium is borne out by this study. (Allen
true pollen tube and that the pollen grain is moved 1946a, p. 555–556)
from the stigmatic flap to the nucellus by a drop
Allen had argued that “the mature pollen
of liquid. The intine is said to be rather
grain is intermediate between Pinus and Abies
homologous differentiated into an outer layer,
with respect to division of the antheridial cell”
which swells readily and an inner very thin layer.
(1943, p. 660), and he largely agreed with
In his discussion of the structure of the male
Lawson (1909) regard- ing the development of
ga- metophyte in gymnosperms, Sterling (1963)
the female gametophyte. In a later paper, Allen
noted that the archegonia of gymnosperms are
(1946b) described the ontology of the
very similar to those of bryophytes, but that the
microsporangium of Pseudotsuga in detail,
same cannot be said for the male gametophyte. He
noting that for this species, “the sporangium is
also noted that “the development of the male
not invested by a true sporophyll epidermis as
gametophyte of Pinus can serve as the type for the
in the case of angiosperms” and that on this
family” (p. 188). Sterling proposed a new
basis “the mi- crosporangium of Pseudotsuga is
nomenclature, following which in Pinus, the
closely homologous with the sporangia of the
embryonic cell gives rise to the antheridial initial,
lower eusporangiate tra- cheophytes.
which in turn gives rise to the generative and tube
Furthermore, there is a close similarity between
cells, the former, to the spermatogenous and the
the non-stratified structure of the vegetative shoot
sterile cells. Finally, the spermatogenous cell
apex and the non-stratified structure of the
produces the two male gametes.
microsporangium” (p. 551).
Ho and Owens (1974) observed the following:
Another paper by Allen (1946a) summarized the
(1) The size of the microstrobili increase acropetally
along the shoot. (2) The average microsporophyll proembryology of Douglas-fir:
contains 64.5 (33–106) micro sporophylls. While An intensive study of several hundred archegonia of
Sziklai reports from
Pseudotsuga has provided evidence that
proembryo formation is essentially similar to that
in Pinus with the
Chapter 6. Flowering 153
the root and the shoot in Pseudotsuga
exception that the completed proembryo of the former
consists of only three tiers.
All four pollen-tube nuclei are discharged into the
egg; three of these, the “supernumerary nuclei,” may
behave in various ways. They may disorganize, they
may fragment “amitotically,” mitoses may take place
often accompanied by cytokinesis, they may “fuse”
with one another and finally degenerate. Apparently
there is no normal behavior and more than three
nuclei are commonly present near the neck region of
the egg. . . .
Although no evidence for a pairing of maternal
and paternal chromosomes at syngamy was obtained,
in the two preparations showing critical stages of the
first metaphase, there was an indication of loose
pairing of chromosomes of similar size. The
suggestion is made that some of the peculiarities of
later embryology in certain conifers may be the result
of pairing of chromosomes at syngamy.
Simple polyembryony occurs in Pseudotsuga and
there is no evidence of cleavage. All four embryo
lineages from the one zygote may contribute equally
to the late embryo, or two lineages may overtop the
others and give rise to the apical initials. The latter
appears to be the more common program. There is
some evidence that occasionally one lineage may
overtop the other three and that its terminal cell may
become the apical cell of the entire embryo. Certain
facts support the concept that simple polyembryony is
less specialized than cleavage although the reverse
viewpoint seems to be generally accepted. It is
concluded that the relation between simple and
cleavage polyembryony is as yet obscure.
Embrogeny has been divided into early and late
stages, the artificial separation being suggested by the
appearance of root generative initials which set apart
the two highly meristematic regions—the stele
promeristem, and the massive rib meristem which is
continuous with the suspensor system. (Allen 1946a,
p. 676)
In yet another paper in this series, Allen
(1947b) discussed the development of the apical
meristems of Douglas-fir:
Arber (1941) has speculated upon the nature of the
angiosperm root and suggested that the shoot may be
likened to a periclinal chimera the inner component of
which is of root nature. She pointed out that the root
has no power of producing either leaves or
sporogenous tis- sue which usually arise from the
more superficial layers of the shoot; that the root may
represent a partial shoot with external
imcompleteness. Arber suggested that the tendency of
the root to divest itself of layers correspond- ing to the
external shoot layers may be significant. Thus,
especially in the dicotyledons, the original cortex is
cast off more or less completely by cork formation.
In view of the fact that the present investigation
has dealt with the origin and development of the root
and shoot apices, and that the behavior of the root
apex is consistent throughout the various stages
examined, it would seem both permissible and
desirable to examine Arber’s hypothesis in the light of
the available evidence. The writer cannot agree with
Arber’s hypothesis because the limited evidence
suggests an entirely different rela- tionship between
(Bloch 1943, pp. 290-293).
and perhaps in many other seed plants. The
(6) Some similarity exists between the zonal pattern of
important points considered are listed in order to
the stelar apex of the root and the zonal pattern of the
present clearly a new viewpoint. . . .
entire shoot apex. The peripheral tissue zone and the
(1) The shoot initials are superficial from the central
beginning; the root initials are always internal.
The former-add cells in one direction only; the
root initials add cells both inwardly to the stele
and outwardly to the mantle. The derivatives of
the shoot apical initials may be ho- mologous
with the inner derivatives of the root initials, the
peripheral mantle of the root apex may have no
homologue in the shoot, and the stele of the
primary root may be homologous with the entire
primary shoot.
(2) The embryonic cortex of the embryo is set off
from the embryonic stele by the appearance of
the root ini- tials and the development which
follows; it is a mantle which completely
surrounds the stele of the hypocotyl and radicle
and, in the dormant embryo, has no con- nection
with the embryonic shoot apex. Outward de-
rivatives of the initials add to the cortex of the
root and the fundamental pattern laid down in
embryogeny is maintained indefinitely. The
cortex arises in the root in close proximity to the
initials. On the other hand, inward derivatives of
the shoot apical initials give rise to the shoot
cortex; development of the latter is retarded and
is possibly related to foliar differentiation or at
least to the provascular differentiation of the leaf
traces (Louis 1935, Kaplan 1937, Barthelmess
1937, Wetmore 1943). There is no region of the
primary root which can be considered
homologous with the cortex of the shoot unless
the root pericycle is cortical in nature. Cortex
and rootcap of the root apices of Pseudotsuga
appear to be two regions of one and the same
mantle, the rootcap being set off from the
remainder by the addition of new cell lineages
from the initials, which displace a part of the
mantle tissue and result in its eventual sloughing.
(3) Lateral appendages originate from the
peripheral tissue of the root stele and not from
the superficial layers of the root itself; the
pericycle is analogous and may be homologous
with the generative layers of the shoot apex.
(4) The first formed phellogen often arises in
the outer cortex of the shoot (Foster 1942, p.
107) but takes origin in the pericycle region of
the root (loc. cit., p. 134). As Arber has
pointed out, there is a general tendency in
woody plants for the superficial layers of the
root to be cast off very early in development.
Once second- ary growth occurs, the structure
of root and shoot is remarkably similar.
(5) Lateral roots originate from the pericycle and
a con- siderable mass of tissue is formed before
there is any differentiation of apical initials. Then
the latter appear deep within the massive
primordium (Reinke 1872, Guttenberg 1941).
The apical meristem of the lateral root has an
origin almost identical with that of the embryo
radicle. It seems likely, therefore, that similar
factors operate which destine the endogenous
primordium to become a root and not a leaf, and
to develop initials which lie deep within the tissue
154 Douglas-fir: The Genus Pseudotsuga
tissue zone of the shoot apex have their counterparts and older sporophyte. With increasing age there is a
in the stele of the root. The peripheral tissue zone of tendency for more rapid differential and maturation of
the root stelar apex produces procambium, pericycle, rootcap tissues and for a more complete contribution
and branch roots. by the column to the peripheral tissue of the rootcap.
(7) There is no “true epidermis” in the root according The embryo shoot apex is simple and undifferenti-
to Strasburger (1872, 1887), Kroemer (1903), Rumpf ated but that of the growing seedling shows a gradual
(1904), Plaut (1910), and Guttenberg (1941). The increase in complexity and specialization. The shoot
homologue of the shoot epidermis may be represented apex of a seedling 3 months old resembles quite
by the endo- dermis, or by the outer cells of the root closely that of an adult plant. The latter has a zonal
stele which, in Pseudotsuga, do not form a definite pattern very like that of Ginkgo or Sequoia.
“layer” because of plastic adjustments and gliding A major theme characteristic of Pseudotsuga and
growth. probably of other gymnosperms is the toti-potentiality
of surface cells. The various appendages of the shoot
(8) The suggestion has been made that the complex and of the reproductive strobili originate from the
tis- sue pattern of the root apex is a result partly of activity of surface cells and all tissues of the shoot
active elongation of the stele and passive expansion and strobilus are traceable back to surface initials.
of the peripheral mantle (Lundegardh 1914). If the This contrasts with the behavior in many angiosperms
root stele is indeed homologous with the whole which exhibit distinctly stratified structures and
shoot, it might be expected that the primary forces organs.
of polarity, whatever their nature, would take effect A recent theory as to the nature of the root is dis-
largely in the central core of the root. cussed and certain evidence presented in favor of an
(9) According to Goebel (1905, pp. 226-227), in the alternative hypothesis. The stele of the primary root is
Pteridophyta and Spermatophyta “there are a number considered homologous with the whole primary shoot;
of cases in which, sometimes regularly, sometimes oc- the peripheral tissues of the root, that is, the “cortex”
casionally, roots become transformed into shoots at and the rhizodermis, have no counterparts in the
the apex by throwing off their rootcaps and forming shoot. (Allen 1947b, p. 210)
leaves.” On page 228, he added, “The transformation
of roots into shoots is, in my opinion, only an In the penultimate paper of the series Allen
individual case of the general phenomenon that shoots (1947a) summarized as follows:
arise upon roots.” Finally, Goebel stated (p. 233) “An The stele promeristem of the embryo at the beginning
actual transforma- tion of a shoot into a root has, as
of the late stage is delimited at its ends by the free
yet, not been shown.” Bower (1908, p. 219) observed:
embryo apex and by the root initials, and is
“It may be found that roots grow on directly into
surrounded by the poorly defined cortex
normal leafy shoots, as in certain Ferns, Aroids, and
promeristem. The stele and cortex promeristems
Orchids, etc.; the converse, however, has not yet been
enlarge mainly by intercalary growth to form the
shown to occur.” Such data are in accord with the
embryonic stele and the embryonic cortex of the
hypothesis here outlined but are hardly explainable on
dormant embryo. The rib meristem, lying between
the basis of Arber’s hypothesis.
the root initials and the suspensor system, adds to
The hypothesis is advanced and may be tested
the embryonic cortex and to the suspensor but
experi- mentally that the primary root of Pseudotsuga is
contributes mainly to the rootcap region of the
homologous with the whole primary shoot, having in
mature embryo. The shoot apex of the embryo
addition an outer mantle of tissue which has no
counterpart in the shoot, that the primary root has the arises from the free apex as a result of the activity
same tendencies and capaci- ties as the primary shoot of surface cells and precedes the cotyleden in
but that these are expressed in different ways, partly appearance. The latter, usually six or seven in
because of the outer mantle and the internal position number, are initiated by the activity of surface cells
of the root initials, and partly because of the unknown in the shoulder of tissue which surrounds the shoot
factors which influence the root and which are apex. During embryogeny there is at no time a
different from those which affect the shoot. A study discrete “dermatogen” or protoderm.
of the transition zone of seedlings might provide The “histogen” concept cannot logically be
further clues as to the homologies between the root applied to the root apical meristem. The latter has
and the shoot. been divided into three mother-cell zones for
Finally, it should be emphasized that toti- descriptive purposes but predetermination is not
potentiality of embryonic surface cells (Schüepp implied. From each zone is derived in large part
1926) is exhibited by Pseudotsuga in their actual respectively the embryonic stele, the embryonic
contribution to many and varied tissues of the plant cortex, and the column, the latter giving rise to the
body, including the sporangia (Allen 1946b). This is bulk of the rootcap. The differentiated dermal
not surprising in view of the fact that all cells of the system of the primary root is termed the
primary shoot may be traced ulti- mately to the “rhizodermis” because of certain fundamental
divisions of superficial cells at the shoot apex. (Allen differences between it and the epidermis of the
1947b, pp. 209-210) shoot, hypocotyl, and cotyle- dons, differences
which were recognized by the botanists of the
Allen (1947b) then summarized as follows: nineteenth century. (Allen 1947a, p. 79)
The apical zonation of the root as laid down during
embryogeny changes very little as the plant grows
The development of the ovule in Douglas-fir is
older and is fundamentally the same in embryo, detailed in these excerpts from Allen (1963, num-
seedling,
bering added):
Chapter 6. Flowering 155
between Pseudotsuga and Pinus insofar as the pollination
1. Morphological differentiation of ovulate buds can mechanism is concerned. (Allen 1947a, p. 393)
usually be recognized in July ... but ovules do not ap-
pear until the following March. Megaspore mother
cells, however, are evident in the late fall—usually by
October and as early in one instance as August [in
Vancouver, British Columbia]. (Allen 1947a, p. 387)
2. The theoretical implications of the timing of
megaspore mother-cell appearance and the
development of the ovule are interesting: the site of
the mother cell foreshadows ovule differentiation—
whether or not it predetermines it. Meiosis occurs
soon after the integument has begun to form from the
enlarging nucellus. (Allen 1947a, p. 387)
3. Growth of the developing ovule and its scale pro-
duce a shift in the orientation of the mother cell which
begins to elongate prior to meiosis. Mother cell and
integument begin to face the outer edge and base of
the scale, the position assumed at pollination. The
adaxial portion of the integument grows much more
than the abaxial to produce the completely one-sided
ovule
tip described by Doyle (1926) and by Doyle and
O’Leary (1935). This stigmatic tip is nearly spherical,
its surface is well supplied with unicellular hairs, and
the crack between the two unequal lips is oriented
away from the cone axis and upward in the erect and
receptive conelet. (Allen 1947a, p. 391)
4. By the time pollination takes place, the integument
tip consists of two unequal “lips” appressed together
to form a closed crack facing upward in the erect
strobilus and toward the interscale cavity. The slit
between the two lobes is, in effect, the closed mouth
of the micropyle. The near spherical tip is well
covered with unicellular hairs, presumably sticky
since pollen grains adhere to them. When receptive,
most of the stigmatus protrudes beyond the edge of
the subtending scale. The tips,
in total, occupy a substantial part of the space
between the scales and form an effective pollen-
catching screen. (Allen 1947a, p. 391)
5. Within a week or so after pollination, the pollen
grains have been effectively trapped and largely
covered as a result of growth of the integument tip.
When pollen is abundant, some or many pollen grains
may be excluded and germinate futilely on the surface
of the neck along with those that adhered to scales or
bracts. The ef-
fective pollen grains are contained within a chamber
which is in direct connection with the relatively
narrow micropylar channel leading to the nucellar
chamber. . . .
The pollen grains remain attached to the stigmatic
hairs and become free only when the young male
gametophyte escapes from its exine. (Allen 1947a, p.
392)
6. In Pinus the pollen is picked up by a pollination
droplet and floats or is drawn into the micropyle to
rest on the surface of the nucellus. In Pseudotsuga, the
pollen remains in the micropylar chamber, held there
initially by the stigmatic hairs; only the pollen tubes
eventually reach the nucellus. In contrast, pollen of
Larix is transferred by fluid to the nucellus after it has
been contained tem- porarily within the integument
tip. In this respect, Larix represents a genus midway
Ovulate strobilus Welwitschia and Gnetum are the only exceptions
(Maheshwari and Singh (1967). As they
The research papers discussing the
explained,
female strobilus and the enclosed ovules
and female gametophyte are not nearly The female gametophytes of the pteridophytes and the
gymnosperms are comparable in the following
as voluminous as that concerning respects:
pollen. Nonetheless, as Florin (1954) (1) both are multicellular; (2) both serve the dual
observed, the ovulate strobilus of function of bearing the archegonia and of nourishing
the young embryo; and (3) the structure of the
conifers has been the subject of a archegonium is es- sentially similar in the two groups;
number of papers during the past two in both cases they have a venter, an egg cell, a ventral
canal cell (not of universal occurrence in
centuries (see previous discussion on gymnosperms) and a variable
flowering). Much of this work has been
concerned with morphological
questions. Chamberlain (1935) argued
as follows:
There can be no doubt that in the
evolution of the ar- chegonium there has
been a gradual reduction in the length of
the neck and in the number of neck canal
cells, which, phylogenetically, are
probably eggs. In the lower Filicales there
are two neck canal cells; in the higher
homosporous forms, only one neck canal
cell with two nuclei; and in the
heterosporous genera, even the mitosis
has failed to take place, so that there is
only one neck canal cell with one nucleus.
In the gymnosperms the mitosis which, in
the pteridophytes, gives rise to the neck
canal and ventral series, is suppressed, so
that the ventral canal mitosis takes place
in the cell which, in Pteris, gives rise to a
primary neck canal cell and a central cell.
(Chamberlain 1935, p. 330)

While Bierhorst (1971) wrote, “Both


mega and micro sporangiate
fructifications are clearly strobiloid in
the Pinaceae unlike members of certain
other co- nifer families in which the
cone like nature of the fructification
may be obscured by reduction and
modification” (p. 433).
According to Maheshwari and Singh
(1967), “the female gametophyte of
gymnosperms is a large and
multicellular structure, and serves the
double func- tion of bearing the gametes
as well as the nourish- ment of the
developing embryo. This is in contrast
to the situation in angiosperms whose
female ga- metophyte is microscopic
and generally eight-nu- cleate with a
single functional gamete. Archegonia,
although almost invariably borne by the
female gametophyte of gymnosperms . .
. are unknown in angiosperms” (p. 88).
156 Douglas-fir: The Genus Pseudotsuga
number of neck cells. However, in the pteridophytes shoots. (Allen and Owens 1972, p. 65)
the gametophyte is usually free-living and green,
whereas in the gymnosperms it is parasitic on the
sporophyte. Further, there are no neck canal cells in
the gymno- sperms. These differences appear to be
related to the di- rection of evolution, which has
resulted in a diminishing capacity of the gametophyte
for independent existence. The similarities, on the
other hand, speak strongly for the homologies in the
structure of the female gametophyte of the two
groups. (Maheshwari and Singh 1967, p. 1967)

More recent investigations have discussed the


physiology of the development of the strobilus
and its enclosed ovules, archegonia, and female
game- tophyte. There are a number of reviews in
this area, including Konar and Oberoi (1969),
Maheshwari and Singh (1966), Allen and Owens
(1972; this excellent paper discusses the research
of the female cone of Douglas-fir until 1972),
Owens and Blake (1986), Pennell (1988), and
Sedgeley and Griffen (1989). Much of the
material discussed in these papers is, perhaps, too
detailed for this book, or is not spe- cific to
Douglas-fir. Accordingly, we will attempt to
synthesize the major points these papers discuss.
Maheshwari and Singh (1967) found that “the
female gametophyte of gymnosperms is a large
mul- ticellular structure, and serves the double
function of bearing gametes as well as the
nourishment of the developing embryo” (p. 88).
Konar and Oberoi (1969) noted that the ovuleaf
Pseudotsuga in unitegmic and crassinucellate and
that the number of archegonia vary from 1 to 7 in
the micropylar end of the game- tophyte. Allen
and Owens (1972) made the following
observations (numbering added):
1. The seed cone is a compound strobilus, in that it
consists of an axis or stem bearing a series of
usually spinally arranged bracts. (Allen and Owens
1972, p. 61)
2. Bracts are initiated over a period of 2½ months,
from mid-July until the end of September, but the rate
of bract initiation is not constant during this time.
Half the final number of bracts are initiated during the
first month . . . (Allen and Owens 1972, p. 65)
3. Ovuliferous-scale initiation begins about the first of
September, 5 months after the seed cone is initiated
and after over half the final number of bracts have
been initiated . . . (Allen and Owens 1972, p. 65)
4. Although, morphologically, ovuliferous scales are
modified lateral shoots (Doak 1935), their initiation
and early development are different from other types
of lat- eral shoots. The ovuliferous scale is more truly
axillary in origin instead of arising from cortical cells
above the axillary region as do vegetative lateral
5. In the latter part of September, megaspore Pennell (1988) argued that, “in comparison
mother cells begin to differentiate and most have
differentiated by mid-October. (Allen and Owens
with the development of the microsporangium the
1972, p. 66) events which take place within the ovules of
conifers are
6. Unlike pollen-cone buds, no evidence of early
stages of meiosis in the fall or the diffuse
diplotene stage has been observed in the
megaspore mother cells. (Allen and Owens 1972,
p. 66)

7. The seed-cone bud shows mitotic activity in


both ovuliferous scales and bracts until early
November at lower elevations, when they
become dormant. (Allen and Owens 1972, p. 66)

8. Development of the ovule resumes about mid-


Febru- ary and coincides with meiosis of both
pollen mother cells and megaspore mother cells.
(Allen and Owens 1972, p. 66)

9. The female gametophyte of Douglas-fir is


not fully developed until late in May, just
before fertilization and almost two months
after pollination. (Allen and Owens 1972, p.
76)

10. In Douglas-fir, four megaspores appear to result from


meiosis of the megaspore mother cell.
development of more than one megaspore has
not been observed. (Allen and Owens 1972, p.
77)

11. Most, if not all, of the superficial cells at the


apex of the archegonium are potentially
archegonial initials even though the number that
fully develops is usually four to six. The most
common number in Douglas-fir is four (Allen
and Owens 1972, pp. 77-80)

According to Owens and Blake (1985, p.


68), “the archegonial jacket, neck cells,
ventral canal cell, and egg constitute an
archegonium.” Owens and Blake (1985) and
Allen and Owens (1972) noted that Douglas
fir is monoecious and that the reproductive
buds and vegetative buds are initiated at the
onset of vegetative bud growth in the spring.
The buds are found primarily on lateral and
lower surfaces of the shoot. The
megasporangiate buds are primarily on the
distal end of the shoot, the microsporangi- ate
buds, proximal. A given bud initial may
follow one of the five pathways, abort,
become latent, be vegetative, or either male or
female.
The bud differentiation as shown by
histochemi- cal tests occurs in early June
while the buds may be identified anatomically
in July, when shoot growth ceases. The
numbers of cones produced in a given year is
a function of the differentiation, not initia-
tion of the buds.
Chapter 6. Flowering 157
specifically with Douglas-fir.
poorly explored by modern technique. (Why does
only one of the four megaspores resulting from
the meiosis of the spore mother cell develop?)
And little is known about differentiation within
the ovule in conifers” (p. 190).
By contrast, Florin (1954) observed that the
ovu- late strobilus of conifers has been the subject
of research for centuries. Much of this work,
however, has been concerned with the anatomy,
not the physi- ology of cones.
In their review of the research concerning the
development of female reproductive structures,
Sedgley and Griffen (1989) noted the following:
The area where the nucellus and integument join and
are attached to the ovuliferous scale is called the
chalaza. A cell within the nucellus enlarges to form
the megaspore mother cell. In most gymnosperms the
ovule does not develop any further than this prior to
pollination.
Meiosis occurs in the megaspore mother cell
around the time of pollination (Owens and Blake
1985). Three of the four products of meiosis
degenerate, and the re- maining megaspore is
generally the cell at the chalazal end of the tetrad.
This megaspore undergoes numerous nuclear
divisions without cell-wall formation resulting in
hundreds of free nuclei in a common cytoplasm.
(Sedgley and Griffen 1989, p. 31)

Owens (1987) studied the effects of cone


inducing treatments, i.e., root pruning and
gibberellin 4/7, upon the apices of Douglas fir. He
found the follow- ing: (1) root pruning did not
affect the initiation of apices but retarded their
development until early July. At this time, the
apices of shoots differentiated into vegetative or
cone buds or became latent. It is hypothesized that
the slight moisture stress occa- sioned by root
pruning may have been a result of reduced, but
statistically, none significant reduction in water.
(2) The GA4/7 treatment did not affect cone bud
initiation or differentiation.

Pollen physiology
Much of the investigations dealing with various
as- pects of coniferous pollen physiology have
been gen- erally reviewed in several publications
(Webber and Painter 1996, Sedgley and Griffen
1989, Owens and Blake 1985, Binder et al. 1974,
Stanley and Linskens 1974). The following is
largely based on discussion in these papers;
however, no attempt will be made to reference the
papers they discuss, save those dealing
Storage sporophytic Douglas-fir.
Generally, pollen stores better at low Viability of pollen
temperatures and low moisture content. A The vigor of a pollen lot has been estimated in
number of papers have investigated factors sev- eral ways:
affecting pollen vitality after storage because
the use of stored pollen is an integral part of
supplemental pollination programs.
Livingston (1964) reported on pollen viability
after 1, 2, and 3 years of storage at several
temperatures and moisture contents, and of
the effects of freeze drying upon subsequent
pollen viability. The re- sults demonstrated
that pollen moisture was the more limiting
factor affecting pollen viability, with optimum
levels below 10%. Freeze drying aided pollen
survival if it followed air drying, but proved
detrimental to pollen that was not dried. The
author hypothesized that the damage at low
temperatures or freeze-drying was caused by
the formation of ice crystals in pollen with
moisture contents greater than 30% (Ching
and Slabaugh 1966). In other trials with
freeze-drying, Livingston and Ching (1967)
found that if freeze-drying were preceded by
air drying and cold, a high level of viability
might be maintained. In contrast, pollen
stored under room temperature and ambient
humidity degenerated within a year. Other
reports (Charpentier and Bonnet-Masimbert
1983, Webber and Bonnet-Masimbert 1993)
demonstrated that rehydrating Douglas-fir
pollen stored for a year at about 4% moisture
content improved germination, while
Mellerowicz and Bonnet-Masimbert (1986)
found that storing pollen at 4% moisture
content damaged the pollen unless it was
rehydrated prior to germination. Dumont-
BéBoux et al. (1999) sug- gested that this
response is caused by “imbibition shock when
[the stored pollen is] put directly onto culture
media” (p. 11). Webber (1995) argued that
pollen should be stored at moisture contents
between 4% and 8%. Copes (1985, 1987)
found that mature pollen at 4% to 7%
moisture content stored success- fully for 1
and 3 years in liquid nitrogen (-196°C),
whereas fertility of pollen stored at 0°C
decreased after 2 or 3 years. These results
demonstrate that the pollen may be
considerably more cold hardy than
158 Douglas-fir: The Genus Pseudotsuga
Germination in vitro. This method involves plac- essential amino acids and of sucrose and glucose in
ing pollen on a growth medium under controlled Douglas-fir pollen. They further noted
physical conditions and recording the percent of
grains that “germinate.” Barner and Christiansen
(1962), Christiansen (1969), and Ho and Sziklai
(1972) argued that Douglas-fir pollen cannot be
germi- nated in vitro. However, Douglas-fir does
not form a pollen tube in vitro (Webber 1987), so
it is diffi- cult to determine what constitutes
“germination.” Traditionally, pollen that
elongated to two or three times its original
diameter was considered to have “germinated”
(Ching and Ching 1976, Shirazi and Muir 1998).
However, this method is most sensitive to assay
conditions (Webber 1995).
Conductance. A technique developed by
Ching and Ching (1976) used measurements of
the conduc- tance of a solution containing pollen;
it was basically a measure of the integrity of
pollen membranes. But this method is very
sensitive to the hydration state of the pollen
(Webber 1995).
Respiration. Binder and Ballantyne (1975) and
Webber and Bonnet-Masimbert (1993) found that
the respiration of pollen is correlated with other
measures of vitality. According to Webber (1995),
“respiration is the least sensitive test to assay
condi- tions and consistently gives the best
indication of pollen fertility potential in Douglas-
fir” (p. 512 ).
The above tests estimate in vitro germination;
the results generally are correlated with estimates
of sound seed resulting from the use of a pollen
lot in standard pollination applications (Binder
and Ballantyne 1975, Ching and Ching 1975,
Webber 1986, Webber and Bonnet-Masimbert
1993). The latter noted, however, that media
effects and pollen hydra- tion effects must be
considered before regressions of in vitro viability
tests against seed set can be made.
Pollen biochemistry. There are few studies of
the biochemistry of Douglas-fir pollen. Binder et
al. (1974, p. 16) observed that studies of the
physiology of pollen date to 1829 and that 8,000
papers were published on the subject in the 50
years prior to 1963; however, neither they nor
Stanley and Linskens (1974) referenced Douglas-
fir in their reviews. Ching and Ching (1976)
presented data detailing the pres- ence of 21
that these compounds leached much more medium until the pollen grains elon- gated. The
readily from dead as opposed to vigorous pollen was then transferred to fresh media
pollen. Ching and Ching (1976) also found supplemented with flavanols (Kaempferol,
that “the total enzyme activity of quercetin and myricetin). This procedure resulted
ribonuclease, amytase, acid phosphatase and in pollen
protease per 10 mg pollen was reduced with
reducing germinability.” The “specific
activity of these enzymes, however, increased
with decreasing pollen viability, indicating a
preferential retention and perhaps activation
of these hydrolases over other enzymes” (p.
520).
But the relationship of these changes to
reduced viability during storage remains
obscure. In an ear- lier paper, Ching and
Ching (1962) reported that 0.76% to 0.89% of
the dry weight of Douglas-fir pollen is fatty
acids, most of it oleic, palmitoleic and linoleic
acids. In another report, Ching and Ching
(1959) found that pollen “germinated” best
under moderate conditions and that such
“germination” is stimulated by gibberellic
acid. There was, however, no evidence of
pollen tubes presented. In other trials (Ching
et al. 1975), the content of adenosine
triphosphate (ATP) in dry pollen was found to
be correlated with germination rate. Finally,
Muren et al. (1979) conducted a metabolic
study of Douglas-fir pollen germination and
found that starch is the sole food reserve; that
it was adequate for four days of growth
without exogenous sugars; that the energy
charge increased during the first 24 hours of
ger- mination; that GA did not affect
germination; and that respiratory rates
remained constant during the first 48 hours,
and then increased fourfold during the last
half of the 4-day period.
As noted previously, in all the foregoing papers,
pollen was considered to have “germinated”
when it elongated to 2-3 times its original
diameter, and it was generally believed that it
was impossible to stimulate growth of a true
pollen tube. Growth of a pollen tube is
essential to the germination in vitro. However,
Dumont-BéBoux and von Aderkas (1997)
demonstrated that, if the phase procedures
were followed, a tree pollen tube would grow
in vitro. Their procedure consisted of
incubating pol- len grains for 7 days on a
modified Brewbaker and Kwack (1963)
Chapter 6. Flowering 159
Copes and Vance (2000) showed that suspension in
tubes analogous to those found in vitro. Although
cool water reduced
Dumont-BéBoux and von Aderkas (1997) found
the two-phase technique to be essential to pollen
tube formation, an earlier paper from the same
labora- tory by Fernando et al. (1997) reported
successful tube growth without flavanols; and, in
a later report, Dumont-BéBoux et al. (1999),
agreed. Aft (1961) had shown that
dehydroquercetin is endogenous to Douglas-fir.
The survey of papers concerned with Douglas-fir
pollen has found no references that discussed how
the results of the above procedure might be
affected by environmental or procedural factors;
however, since the first step involved the
elongation of pollen grains, it is assumed that the
procedure by Dumont-BéBoux and vonArderkas
(1997) would have been affected in a manner
similar to that discussed for earlier in vitro trials.
Other studies discussing the effects of various
procedures upon pollen viability (as measured by
elongation or staining, rather than pollen tube for-
mation) include (1) Livingston and Stetller (1973),
who found that gamma radiation speeded pollen
elongations (possibly causing increased
metabolism); and (2) Shirazi and Muir (1998, p.
341), who reported that “formaldehyde at target
concentrations of 300, 600, 900 and 1200mmol
m–3 reduced germination of Douglas fir
[Pseudotsuga menziesii (Mirbel) Franco.] pollen in
vitro,” as measured by pollen grain elonga- tion.
Shirazi and Muir (1998, p. 341) also presented
evidence that live Douglas-fir pollen, as shown by
2–3-5-triphenyltetrazolium chloride (TTC) staining,
had the capacity to significantly reduce levels of
formaldehyde in the germination media,
“potentially attributable to uptake by pollen or a
detoxification mechanism.”
Dumont-BéBoux et al. (1999) reviewed studies
demonstrating that moisture content may have a
significant affect on pollen growth, particularly
in the case of dry stored pollen; rapid
rehydration may damage membranes. They also
noted that the major effect of PGE (
prostaglandin E) is as an osmoticum. The
mechanism is complex, but this material appar-
ently acts to stabilize membranes. Other
materials shown to be important to pollen
growth include Ca(NO3)2 and H3BO3 (Brewbaker
and Kwack 1963, Fernando et al. 1997). Finally,
slow at first; archegonia start to form in early
pollen viability by about 3% per day, and that frost May, (there are commonly archegonia in
damage to pollen in the spring was not uncommon. Douglas-fir) and are mature in late May, just
Fertilization before fertiliza- tion when the female
gametophte is mature.
Pollination and fertilization in Douglas-fir
represent a series of events that terminate in
the fertilization of the egg cell by male
gametes released from the pollen tube. Owens
and Blake (1985, p. 71) state that “discussions
of fertilization may include all stages from
pollen structure through gamete fusion (syn-
gamy).” We will limit our discussion to the
events including the arrival of pollen on the
stigma until the fusion of male and female
gametes.
Fertilization in conifers has been the
subject of research for about a century.
However, as Allen and Owens (1972, p. 101)
noted, “most of the information available
(refertilization in conifers) is based on very
early work, which, though carefully done, did
not have the benefit of our modern, refined
equipment and techniques.”
In an excellent, detailed discussion, Allen
and Owens (1972) gave the results of their
research and reviewed the previous papers
concerned with pol- lination and fertilization
of Douglas-fir. Their pre- sentation is too
technical and detailed for this book, but we
strongly recommend that readers concerned
with all the numerous events inherent in the
repro- ductive physiology of this plant consult
this excellent monograph. We will confine our
coverage to what appear to be their major
points.
1. Bracts are first initiated in mid-summer,
and this initiation continues until
dormancy in mid- September. Growth
resumes in early spring so that the mature
bract functions as a funnel to guide
pollen grains to the stigmatic tips.
2. The stigmatic tips are unequal and are
part of the integument of the ovule; the
adaxial part of the integument forms a
time covered with fine, sticky hairs.
After the pollen grains are trapped on the
stigmatic surface, the two stigmatic tips
grow together, sealing the pollen grains
at the mouth of the micropyle.
3. Development of the female gametophyte is
160 Douglas-fir: The Genus Pseudotsuga
4. Pollen germination: neck cells of the archegonium and then releases, through
Pollen need not be in the micropyle to germinate but
full normal development would probably not occur
on the surface of the bract or ovuliferous scale. Pollen
germinates within the micropyle about three weeks
after it is engulfed by the stigmatic tip of the
integument. Pollen grains are not in contact with the
nucellus when they germinate; rather, germination
occurs when pol- len is still adhering to the stigmatic
tip just inside the micropylar canal and a considerable
distance (several hundred microns) from the nucellus.
(Allen and
Owens 1972, p. 91)

5. There is no pollination drop in Douglas-fir:


The absence of a pollination drop means that the
pollen grain has to grow (elongate) inward to the tip
of the nucellus. . . .
At germination, the pollen grain swells and the
exine of the spore wall splits open. There is no
apparent pore or line along which the exine splits. The
intine remains intact and forms the very plastic wall
as the pollen grain elongates. Normally, the distal end
of the pollen grain (opposite the prothallial cells)
forms the advancing tip of the elongating pollen grain.
The proximal end usually remains within the broken
exine, or the exine is shed entirely. In either case, no
distinct pollen tube forms at this time but the entire
pollen grain forms a long tubular structure. The
intine is very plastic and capable of
considerable extension. This could be attributed to the
nature of the intine which, in Douglas-fir, consists of
cellulose and pectin and possibly callose, as in other
plants........although tests for the latter were not made
in Douglas-fir........The thick intine becomes very thin
during pollen elongation....which continued for
several
weeks. (Allen and Owens 1972, p. 91)
6. Fertilization:
The term fertilization as applied to plants may
involve more than the fusion of a male gamete with
an egg. In gymnosperms, fertilization may involve
several other and often unusual events from the
time the pollen tube reaches the neck of the
archegonium until the first di- vision of the zygote.
The fate of all nuclei and cells passing through the
pollen tube of Douglas-fir must be considered,
since they do not all disappear after fusion of the
male gamete and the egg. The misinterpretation of
these structures once they are within the egg has led
to erroneous conclusions regarding the normal
pattern of fertilization and early embryo
development. . . .
Fertilization in Douglas-fir normally occurs
between June 1 and June 20 at lower elevations in the
Pacific Northwest and British Columbia. In any one
year there is a variation of several days between
ovules on an individual tree. Certain trees tend to be
much earlier or later than the average and exhibit this
behavior year after year. . . .
The female gametophyte is fully developed at the
time of fertilization. (Allen and Owens 1972, p. 97)

Additionally, they noted the following:


The pollen tube penetrates the nucellar tissue and the
its tip, its entire contents into the egg cell; Columbia.
the tube nucleus, the two male gametes
and the stalk cell. . . .
Owens et al. (1981) reported that seed cones
Details of this process are difficult to re- mained most receptive to pollen for at least 4
observe because of the disruption of days, but that 6 days after the conelets became
tissues during pollen-tube penetration.
The larger male gamete moves rapidly receptive, the stigmatic tips showed less
toward the egg nucleus where nuclear receptivity, and that by 10 days, the entrance to
fusion occurs. The remaining the micropyle was closed. They recommended that
supernumerary nuclei are often left
close to the neck or somewhere between the best time for pollina- tion to occur was 4 days
the neck and the fusion nucleus . . . after the seed cones were half out of the scales;
Several fates are possible for the supernumerary nuclei: that the optimum number of pollen grains per
(1) they may disorganize during a period of several days;
(2) they may fragment and thus increase stigmatic tip was 11; and that the optimum
the number of apparent free nuclei in the number of pollen grains within each my-
neck region; (3) one or more of them may
divide mitotically and cell walls may
form in the next region, or (4) the nuclei
may fuse to form
larger nuclei which gradually degenerate.
The ventral canal nucleus may fuse
with supernumer- ary nuclei or undergo
independent division within its own cell
wall. This may appear similar to triple
fusion
in angiosperms, where the endosperm
tissue results. The significance of
fusion, division or fragmentation of
supernumerary and ventral canal nuclei
is not known but their presence can
result in the misinterpretation of the
normal course of fertilization and early
embryo development. It should be
recognized that these do
not represent either the endosperm of
angiosperms or a second embryo.
Fusion of male and female nuclei in
Douglas-fir is similar to that described
for other members of the Pinaceae.
The egg nucleus flattens somewhat on
the side nearest the approaching male
gamete, as in Pinus. The male gamete
meets and gradually sinks into the egg
nucleus but retains its identity for some
time.
. . . The membranes around both the male
and female nuclei soon disappear, the two
groups of chromosomes become evident
and spindle fibers become visibly
associated with each group of chromosomes. The
two spindle figures come together
laterally to form a common, multi-polar
spindle with its main axis usually
perpendicular to the long axis of the egg.
The multi-polar spindle figure appears to
contract and become bipolar by
metaphase of the first division of the
zygote. (Allen
and Owens 1972, p. 100)

A significant amount of research


concerning various aspects of
fertilization in Douglas-fir has benefited
from improved technology, much of it
by scientists working in Professor
Owen’s laboratory in Victoria, British
Chapter 6. Flowering 161
is maternal. The mechanism by which mtDNA, thus
cropylar canal was 3. In a later report (Owens and mitochondrial characters, are maternally inherited is by the
aggregation of mitochondria in the perinuclear
Simpson 1982), it was noted that pollen applied to
female cones 3 to 5 days after they became recep-
tive was taken into the micropyle more than
pollen applied at later dates. They further noted
that the seeds produced after an average of 1.2
pollen grains per micropyle were as viable as
those resulting from 3.4 or more grains per
mycropyle. Webber and Painter (1996) reviewed
substantial data demontrat- ing that first arriving
pollen has a distinct advantage in subsequent
fertilization. Knox and Singh (1987) observed that
for angiosperms, studies of the sig- nificance of
pollen load on subsequent fertilization have just
begun.
Owens and Morris (1988) presented the results
of an ultrastructural study designed to investigate
the mode of inheritance of mitochondria in
Douglas-fir; some of their observations follow:
Pollination occurred in April and was followed by 6
weeks when pollen elongated within the micropylar
canal. A pollen tube then formed, penetrated the
nucel- lus and fertilization occurred by mid-June
Embryos
developed over the next 2 months.
The engulfed pollen swelled, ruptured the exine
and elongated. The nucellus tip formed a minute
secretion stimulating pollen-tube formation. A narrow
pollen tube penetrated between the loose outer
nucellar cells. Deeper in the nucellus, cells which
came in contact with the tip collapsed. Two to four
pollen tubes commonly penetrated each nucellus.
During pollen-tube growth the tube nucleus remained
near the tip of the pollen tube followed by the large
body cell and small stalk cell. . . .
Pollen tubes grew to the distal end of the
megagame- tophyte, penetrated the megaspore wall
then grew into one of the archegonial chambers found
above each group of neck cells. The body cell
settled into this pocket
and divided to form the two male gametes. (Owens
and Morris 1988, pp. 339-340)
They concluded,
In Douglas-fir and perhaps other conifers, the
mechanism by which cpDNA, thus plastid-
associated characters, are paternally inherited, is by
compartmentalisation of plastids in the body cell
and exclusion and destruc- tion of maternal plastids
during egg development. Our ultrastructural
observations of Douglas-fir and earlier reports for
Pinus (Camefort 1962), Larix (Camefort 1967),
Douglas-fir and Chamaecyparis (Chesnoy 1973)
agree with the pattern of strictly paternal cpDNA
inheritance reported by Neale et al. (1986) using
restriction fragment length polymorphism (RFLP)
techniques on Douglas- fir and other members of
the Pinaceae, Cupressaceae and Taxodiaceae.
Preliminary studies using RFLP tech- niques in
conifers indicate that mtDNA inheritance
In a series of reports, Takaso and Owens
zone of the egg. However, there is some
paternal con- tribution resulting from the
(1994, 1996), Takaso et al. (1996), and
compartmentalization of paternal organelles vonArderkas and Leary (1999) discussed ovular
which migrate as a cluster with the secretions and their effects on pollen tubes.
neocytoplasm. Ours and earlier ultrastructural
studies suggest that mitochondria are primarily Takaso and Owens (1994, p.
of maternal ori- gin with some contribution
from the cluster of paternal organelles. These
observations suggest further studies of
mitochondrial inheritance are needed using
RFLP techniques. (Owens and Morris 1988, p.
342)

In a subsequent report, Owens and Morris


(1990) verified the above. In yet another
paper concerned with mechanisms of
fertilization, Owens and Morris (1991) made
the following observations:
1. The body-cell (of the pollen) did not
divide until the pollen tube reached either
megaspore mem- brane or entered the
archegonium. At this point it formed the
male gametes (p. 1514).
2. No pollen entered more than one
archegonium, and only one tube was
found on an archegonium (p. 1517).
3. The male nucleus enters the egg cell (p. 1518).
4. The male cluster of organelles does not
mingle with maternal cytoplasm (p. 1519).
5. Male gametes are non-mobile (p. 1923).
6. “Future and more complete RFLP and
ultrastruc- tural studies will reveal the true
variation that exists in the mode of
cytoplasmic inheritance in all conifer
families” (p. 1926).
Crook and Friedman (1992) suggested that
in- creased embryo competition (as evidenced
by simple polyembryony) results in better
adapted germinates than supplemental
pollination (which results in more pollen
tubes per ovule) and the use of clones with
higher levels of archegonia per ovule—both
of which increase embryo competition—
should result in more vigorous germinates.
The detailed study by Takaso and Owens
(1995) on the movement of pollen within the
cone of Douglas-fir largely confirms earlier
work by Owens (1973, 1981). They noted that
the curved bract base effectively guides pollen
to the ovular apex, but that the pollen does not
rest on the surface of either the bract or the
ovuliferous scales (1995, p. 437).
162 Douglas-fir: The Genus Pseudotsuga
504), reported that the morphology of the outer in monticola with megagametophytes dissected from
line and of the plasma membrane of the pollen ap- cones of other genera—Pseudotsuga menziesii,
peared to be affected by secretions from the ovule. Larix
They noted further that pollen tube formation may × eurolepis (now Larix × marschlinsii Coaz) and
be stimulated by material from the ovule. The api- Pinus monticola—and demonstrated no signals
cal degeneration of the occurred before pollen between pollen and megagametophytes.
tube formation (Takaso and Owens 1994, p. 512).
Takaso and Owens (1996) presented detailed Pollen distribution
evi- dence for the presence of three major According to Stanley and Kirby (1973),
secretions in the mycropyle during the Determining quantity, shedding time, and dispersal
fertilization period. They also noted that patterns of pollen is important for ecologists, plant
archegonia formed 4 weeks prior to fertilization, breed- ers, fruit orchard growers, allergists,
palynologists, and farmers. Ecologists are interested
and that cells in the proximal portion of the in the range of viable pollen flight as a contributing
nucellus started to degenerate 5 weeks before factor to interspecific variation; plant breeders are
fertilization and ceased 1 week prior to concerned with the dis- tance necessary to isolate seed
production plots from contaminating pollens. Orchard
fertilization (p. 151). The three secretions were managers must consider factors such as probable
1. an aqueous secretion from the ovular integu- pollen mix and the number of male trees or pollinator
varieties required to pollinate dioecious female trees
ment, which appears to result from cellular of compatible varieties. Allergists evaluate air-borne
breakdown in the apical part of the nucellus pollens capable of inducing allergies in humans;
2. a secretion originating in the female gameto- palynologists analyze the distributions and selective
survival patterns of pollens in ground and water
phyte, which either directly or indirectly sources. (Stanley and Kirby 1973, p. 303)
stimu- lates pollen growth
3. secretion from egg and prothallial cells, which There have been many papers in which the move-
“is liberated to begin and reach to the ment of pollen and the weather factors—such as
micropylar canal when egg cells are ready to tem- perature, wind movement, and rain—that
accept male gametes and induces the have been shown to affect the movement of
formation of pollen tubes” (p. 157). anemophilous pol- len have been discussed.
Takaso et al. (1996, p. 1214) confirmed much of Generally, it is recognized that the sedimentation
the above. Homogenates of megagametophyte, factor, which is determined by the shape, density,
but not the nucellus or integument, stimulated and volume, strongly affects the distance pollen
pollen activity. may move. Schwendemann et al. (2007)
Von Aderkas and Leary (1998, p. 356) found developed a computational model based on
that drops of secretion occur in the micropylar structural characters of pollen grains to investigate
canal during the 2 weeks of late central cell and pollen flight in different conifers, quantitatively
egg cell development, and that the total volume of demonstrating the adaptive significance of sacci
material produced was 1-4 times the micropylar for the aerodynamics of wind pollination.
volume. They speculated that such material Topography also plays a role: Silen and Copes
(which is post- pollination phenomenom) may (1972) reported that contaminating Douglas-fir
play a role in pre- fertilization events including pol- len may move several miles in level areas,
male secretion (p. 356). The material thus far but that this effect was greatly enhanced in narrow
discussed concerns in vitro fertilization. Fernando valleys. In another study, Silen (1962a) found that
et al. (1997, 1998) reported repeated trials of in the great- est amounts of Douglas-fir pollen fell
vitro fertilization, resulting in the first known in within 17 m of the tree, but then 40% as much
vitro fertilization of a coni- fer. Finally, Dumont- pollen was found 610 m from the source, and that
BéBoux et al. (1998), utilizing the methodology a significant amount of pollen moved 2.4 km.
developed in the above studies, showed in vitro Wright (1952) found that nearly all of the
fertilization between genera, co- culturing Larix Douglas-fir pollen was found within 61 m of the
occidentalis, Picea sitchensis, and Pinus source tree, but noted that the sample size was
small and that the tree was short; he esti- mates
that the error could be 2 or 3 times the result.
Chapter 6. Flowering 163
were not
In a second paper (1953), he estimated the
standard deviation of pollen dispersion distance as
18 m; whereas Prat (1995) found the effective
distance of pollen dispersion in a Douglas-fir seed
orchard to be 20–30 m and depended on the wind.
Potter and Rowley (1960) reviewed a large
number of stud- ies reporting pollen distances of
160–100 km, and Douglas-fir pollen at 320 km.
Similarly, Wodehouse (1935) and Erdtman (1943)
noted reports of pollen movement in the hundreds
of kilometers. Ebell and Schmidt (1964) reported
that pollen in bulk moved over 1.6 km (but noted
that the experimental site was particularly windy)
and that it could move upslope from 305 m to
1,219 m. Webber and Painter (1996) noted that
the “dispersal of pollen from a source tree
depends on several factors including pollen size,
morphology, sedimentation velocity, wind
properties (i.e., turbulent flow and convection
currents), and meteorologic conditions (especially
temperature and humidity)” (p. 52).
Lanner (1966) argued, however, that reports in-
volving measurements of the lateral movement of
pollen by wind do not consider the great
movement of pollen by air “shells” and hence do
not provide a true measure of pollen movement.
He reviewed a number of papers that reported
pollen movement in miles, not feet.
Unfortunately, there have been no well-designed
studies to evaluate his hypothesis since its
publication.
Sorensen (1972) presented an approach
whereby pollen from female flowers with an
albino strain could be used to estimate effective
pollen move- ment in a seed orchard, utilizing a
female tree as a receptor. He demonstrated how
such a scheme could measure local pollen
movement and estimate the bulk-distant
movement discussed by Lanner (1966). Adams et
al. (1992) expanded Sorensen’s approach utilizing
genetic characteristics of pollen grains and female
trees to estimate effective polli- nation distances.
Although they found that pollen may move 0.5–
2.0 km, most of the effective pollen originated
from trees within two ranks of the female tree,
and that no pattern of pollen movement could be
detected at distances of less than 30 m from the
mother tree. In contrast to the foregoing, Lowe
and Wheeler (1993), working primarily with
southern pines, showed that 122-m isolation zones
collection efficiency;
sufficient to prevent significant contamination (5) pollen release timed within both the season and the
and suggested that “whenever it is possible, day to maximize the possibility of pollen capture by
receptive conspecifics downwind; (6) relatively close
orchards should be established outside the spacing of compatible plants; (7) vegetational
species range, in areas where the species is structure
scarce, or areas where the phenological
overlap of orchard and surrounding
populations is nil” (p. 51).
Both Bramlett (1981) and Squillace and
Long (1981) reviewed a number of papers
confirming that local pollen distribution
patterns are characterized by a rapid
diminution of pollen grain numbers with
distance from a source tree. Boyer (1966)
suggested that reduction in pollen from a
source tree occurred primarily through
dilution for longleaf pine.

Pollination
Earlier we discussed the development of the
mi- crosporangiate strobilis prior to the
formation and release of pollen. Pennell
(1988) noted that for coni- fers generally,
“once meiosis has been completed a complex
series of events occurs within the cytoplasm
of the spores, and these are ultimately
responsible for the patterning of the
sporoderm” (p. 185). Southwork (1988) noted
that when pollen grains hydrate, the exine,
which is largely sporopollen, separates from
the intine, the inner wall of the pollen grain.
Fechner (1978) stated that increasing
temperatures and reduc- ing humidity
generally stimulate the development of
coniferous “flowers” and that, in wind-
pollinated species, the flowers open and shed
pollen during warm, dry periods. Conversely,
low temperatures and rainy weather retard
these events; Owens (1982) agrees for
Douglas-fir.
Douglas-fir is a wind-pollinated species.
We dis- cussed the vagaries of pollen
movement in another section. Whitehead
(1983) discussed the various factors affecting
the success of wind pollination as follows:
Wind pollination is most likely to be successful if
cer- tain idealized conditions are met. These
include (1) the production of large numbers of
pollen grains; (2) pollen grains with appropriate
aerodynamic characteristics; (3) flower and
inflorescence structure and location on the plant
designed to maximize the probability of pollen’s
entrainment in moving air; (4) stigmatic surfaces
struc- tured and positioned to maximize
164 Douglas-fir: The Genus Pseudotsuga
that is relatively open to minimize filtration of pollen Dowding also stated that wind pollination is both a
by nonstigmatic surfaces; (8) wind velocity within an
accept- able range to ensure transport and minimize “primitive” in gymnosperms and an “advanced”
downwind dispersion; (9) relatively low humidity and characteristic in angiosperms (p. 421). He justified the
a low prob- ability of rainfall; and (10) unambiguous
environmental cues to coordinate flowering.
(Whitehead 1983, p. 98)

Many of these factors have not been evaluated


spe- cifically for Douglas-fir. However, given data
such as those reported by Silen (1962a),
Whitehead further noted that relatively large
pollen grains, like those of Douglas-fir, have a
better chance of being cap- tured by stigmatic
surfaces than do lighter grains. Whitehead also
found that “the frequency of wind pollination
increases with elevation in mountainous regions,
both in temperate and tropical latitudes” (p. 103).
He noted that “anemophily is much more
common in temperate and boreal zones, than it is
in tropical climates,” and that “anemophily is rare
in extremely arid environments” (p. 106).
Regal (1982) suggested that the
preponderance of animal pollination may be
correlated with arid- ity, possibly because of
short climatic uncertainty. Dowding (1987, pp.
421–22) compiled the following advantages and
disadvantages of anemophily:
The advantages of wind pollination are:
1. Lack of dependence on an animal agent for
pol- lination and seed set. This is of particular
advantage in higher latitudes where peak insect
populations are reached in late summer, more than
halfway through the growing season. Wind-
pollinated flowers often open before the growing
season starts.
2. Male gametes by consumption of pollen as food
by the vector.
3. Much lower capital and maintenance
expenditure on showy non-photosynthetic petals and
on nectaries.
4. The possibility of very long distance dispersal
(Erdtmann 1938, Potter and Rowley 1960), though . .
. the probability of successful pollination at distances
greater than 10 km from the source is very small
indeed.
The disadvantages of wind pollination are:
1. Dependence on particular weather patterns for
successful dispersal and deposition . . . (Andersen 1980).
2. Wastage by sedimentation and by impaction
onto nonreceptive surfaces, through washout by rain,
and by premature death caused by UV irradiation and
pos- sibly drying.
3. A very small chance of multiple cross-
fertilization by many pollen grains on each stigma,
hence ovaries usually have single ovules (Corner
1964).
latter by noting that wind-pollinated angiosperms but the results were the same for all species
have evolved from insect pollinated plants. observed, so it may be that the same relationship
Regal (1982) argued that “pollen rain at holds for Douglas-fir.
one point can result from many individuals
even hundreds of miles away. But because the
pollen rain from a single amenophilous tree
delutes rapidly (prob- ably as the inverse of
the square of the distance). The individual’s
probability of fertilization is much reduced by
distances of 6-30 m” (p. 505).
In discussing a relatively rare researched
phe- nomenon, Stephenson and Borsin (1983)
reported that “plants may compete in terms of
male repro- ductive output either before or
after pollination. Prepollination competition
seems probable in many plant populations, but
an unequivocal demonstra- tion of
competition which would involve monitoring
male success before and after removal of
certain donors, has not been made” (p. 124).
Working with pollen from trees with low
and high self fertility, Nakamura and Wheeler
(1992) found that male reproductive success
is not related to degree of self-fertility;
whereas Apsit et al. (1989) demonstrated that
differential male success is appar- ently
genetically determined and that there is male-
female complementarity in Douglas-fir.
Willson and Burley (1983) argued that male
reproductive success is a function of the
quantity of pollen produced and the number of
matings (p. 45).
Webber and Painter (1996) surveyed a
number of reports indicating that for Douglas-
fir, the ma- jor factor affecting pollen success
in fertilizing the ovule, without regard to
pollen quality, is the order in which the pollen
grains arrive at the stigmatic tip; the first
arriving grains apparently occupy the most
favorable sites and, hence, are most successful
(pp. 43–44).
A series of unique reports by Niklas (1984,
1985ab, 1987) and Niklas and Paw U (1982,
1983) discussed the aerodynamics of the
pollen grains of a range of conifers and the air
flow patterns around ovulate cones of the
same species, concluding that the joint effect
of these two factors is to ensure that the pol-
len of a given species will arrive preferentially
at the stigmatic type of the same species. The
species studied did not include Douglas-fir,
7. Seeds
Denis P. Lavender

A
true seed has been described as “A fertilized
mature ovule that possesses an embryonic plant, Seed Dormancy
stored food and a protective coat or coats. The An increasing number of research efforts have
embryo is made up of one or more cot- yledons, a dealt with dormancy in seeds (the seed is an entire
plumule (embryonic bud), hypocotyl (stem or- ganism, well adapted to experimental
portion), and a radicle (rudimentary root)” manipula- tion and, as such, a favored subject for
(Kozlowski and Gunn 1972, p. 5). The major dif- a range of experimental botanists, including
ferences between seeds of angiosperms and that physiologists, biochemists, geneticists, and
of gymnosperms such as Douglas-fir are that the ecologists). As a result, the number and
former are the product of double fertilization complexity of reports dealing with seed dormancy
within an ovary, whereas gymnosperm seeds has expanded greatly during the past half century.
result from a single fertilization and are Vegis (1964, p. 197) suggested that seed
“naked”—not enclosed in an ovary. The vast dormancy has phases, including true dormancy
volume of seed research has focused primarily on and conditional dormancy. These ideas were
the 250,000 species of angio- sperms, which are expanded upon, particularly for seeds, and
tremendously important to the human food discussed in a comprehensive review by Baskin
supply. At less than 1,000 species, gym- nosperms and Baskin (1998). Although much of the vast
have received far less research attention. Research volume of references they discussed are
on Douglas-fir seed has often been concerned concerned with angiospermous seeds, there is
with either seeds that are dormant or are information relevant to Douglas-fir or other
actively germinating. The classic definition of conifers in pages 27–39. They discussed the six
dormancy is “any case in which a tissue broad classes of seed dormancy suggested by
predisposed to elongate does not do so” Nikolaeva (1969): physical, morphological, morpho-
(Doorenbos 1953, p. 1). Kozlowksi and Gunn physiological, chemical, mechanical, and physical
(1972, p. 16) noted that seed dormancy is the plus physiological dormancy, and later noted that
resting stage of seed embryos between Douglas-fir has physiological dormancy, which is
development and germination. The other major a function of the embryo (p. 354).
physiological state of seeds, germination, also as Although the basic physiology of the dormancy
defined by Koller (1972, p. 14) is “a process of seeds is not known, its several manifestations
which starts with the supply of liquid water to the have resulted in the following terms:
dry seed and ends when the growth of the • embryo dormancy—dormancy is seated in the
seedling starts, most commonly by protrusion of embryo itself
the embry- onic radicle through the seed coat.” As • seed-coat dormancy—dormancy is caused by
Bewley and Black (1994) and Bewley (1997) noted, some feature of covering tissues that limit the
however, these definitions do not provide a access of the embryo to environmental
physiological basis for dormancy or germination, factors essential to germination
the mechanisms of which • para dormancy—imposed dormancy,
caused by an environment unfavorable to
are not fully understood.
growth
165
166 Douglas-fir: The Genus Pseudotsuga

• ecto dormancy—innate dormancy, dormancy layers of sand and seeds. The seeds would then be retrieved after
exposure to low temperatures during winter. Stratification will be
seated in the seed discussed more fully in a subsequent section.

• induced secondary dormancy, seeds not


originally dormant but forced into
dormancy by the environment
• relative, conditional dormancy—growth
possible only in a narrow range of
environments
• primary dormancy—dispersed from plant in
a dormant state
• summer predormancy
Baskin and Baskin (1998, p. 29) noted that
“physi- ological dormancy is caused by a
physiological inhibiting mechanism of the embryo
that prevents radicle emergence. Nikolaeva (1977)
distinguished three levels of physiological
dormancy:
Non-deep physiological dormancy is frequent
in weeds, vegetables, flowers and is broken by
exposure to high, not low temperature.
Intermediate physiological dormancy is found
in angiospermous tree seeds and may be broken
by extended stratification.1 GA may substitute for
cold. Deep physiological dormancy, the only
treatment which breaks this dormancy is an
extended chill-
ing period.
The authors do not specify the type of
physiologi- cal dormancy Douglas-fir has, but
other works sug- gest that it is intermediate.
Bewley and Black (1994
p. 201), Bradbeer (1988 p. 39), and Villiers (1972,
p.
224) all agreed that there are at least two general
types of seed dormancy: (1) dormancy seated in
the embryo (physiological) and (2) dormancy
caused in such a manner by the seed coat.
However, they were all concerned with
angiosperms, so it is dif- ficult to assign a definite
dormancy to Douglas-fir on the basis of their
classifications. Seeley (1994), also working with
angiosperms, noted that “in most cases, we have
not adequately determined whether a treatment
really breaks endodormancy (embryo dormancy).
(The controlling mechanism of which we do not
know) or does the treatment promote growth by
some other mechanism also unknown” (p. 615).
The term “stratification” originally derives from the historic farming
practice of digging a pit in the fall and filling it with alternating
Ching and Ching (1973), working despite very detailed analyses, they failed to
primarily with gymnosperms, suggested four define the physiology ba- sic to seed dormancy.
mechanisms by which the seed coat may Bewley (1997) noted that, “it’s worthwhile
cause dormancy: impermeability to water, pondering why so little progress has been made
mechanical resistance and inhibitors in seed toward understanding dormancy.
coat, and low permeability of seed coats to
gases. They also suggested that seed
dormancy may result from immaturity of the
embryo or endogenous dor- mancy of the
embryo. Villiers (1972) largely agreed with
Ching and Ching (1973) that many species
have this last type of dormancy, probably
including Douglas-fir. These species can be
divided into two groups: the positively
photoblastic and those whose dormancy is
broken by extended exposure to low
temperatures, with Douglas-fir and the pines
fall- ing into this last group. Although this
requirement is true for all Douglas-fir, it
apparently varies with ecotypes. For example,
Allen (1960a) reported that coastal Douglas-
fir required longer periods of strati- fication
than did the interior variety. A rationale for
this difference is that the requirement for
stratifica- tion reflected the fact that dormancy
protected the germinates against winter frost
damage and that the erratic nature of spring
frosts in coastal areas neces- sitated a longer
period of protection. Fowler and Dwight
(1964) and Mergen (1963) reported similar
findings for eastern white pine (Pinus strobos)
seed, but Olson and Nienstadt (1959) did not
find this to be true for eastern hemlock
(Tsuga canadensis). Powell (1987a)
hypothesized that stratification is the
equivalent of a “chilling requirement” for
buds and that plants with long chilling
requirements have long stratification periods.
The length of the chilling period required to
stimulate germination has been used as a
measure of the depth of seed dormancy. Seely
(1997) was critical of this interpretation, argu-
ing that “germination is a measure of growth,
not dormancy” (p. 615). Bewley and Black
(1994) pre- sented a long, detailed discussion
of seed dormancy, and concluded that
dormancy represents a block to processes
essential to germination, that it may be
affected by environmental factors such as
light and temperature, and that a genetic
component is definitely in its control. But,
Chapter 7. Seeds 167
Researchers must con- centrate on the very earliest stages
Undoubtedly, one contributing factor is that we of germination, i.e.,
do not know the defining events in germination”
(p. 1056).
Wareing (1965) defined dormancy “as the state
in which germination of the seed is in some way
prevented, even though external conditions are ap-
parently favorable” (p. 103). Bradbeer (1988)
noted that dormancy mechanisms may occur in
two gen- eralities, embryo coverings and the
embryo. In their encyclopedic review of seed
germination and dor- mancy, Baskin and Baskin
(1998, p. 27) noted that two general causes of
dormancy exist: the environ- ment and the seed
itself, and that Douglas-fir has physiological
dormancy: “physiological dormancy is caused by
a physiologically inhibiting mechanism of the
embryo that prevents radicle emergence.
Structures that cover the embryo, including endo-
sperm, seed coats, and inhibitors in walls may
play a role in preventing germination” (p. 29).
They noted that about half of the tree species in
moist, warm temperature woodlands have a non-
dormant seed (p. 352), and that “in the
temperate/arctic vegetation types in which trees
occur, the proportion of trees with non-dormant
seeds decreases with decreases in precipitation”
(p. 562). Corbineau et al. (2002,
p. 315) stated that Douglas-fir has an embryo dor-
mancy, but Bianco et al. (1997, p. 117) reported
that the dormancy is a function of the seed coat.
Bewley and Black (1994) noted “that dormancy
confers the advantage that because seeds of a
given seed crop are differentially dormant, their
germination is tem- porally diverse, which means
that seeds are faced with environments that are
differentially favorable to growth and survival of
germinants” (pp. 199–200). Although they
discussed a wide range of factors associated with
seed dormancy, however, their ex- amples are
overwhelmingly angiosperms.
Biochemistry and Seed Dormancy
Breaking
Seed scientists often believe that the breaking of
seed dormancy and germination are separate
events. For example, according to Leadem (1987):
The breaking of dormancy and the initiation of germi-
nation are two separate and distinct processes, yet it is
apparent from reading the literature that this
somewhat obvious fact is often overlooked.
but they did not identify the genes. Jarvis et al.
during dormancy release, if they are to learn how
PGRs regulate tree seed dormancy. Little
(1997) reported that “late embryogenesis
attention has been given as well to seed abundant protein encoding genes increased in
maturation, but in many tree seeds the induction Douglas-fir seed after one week of stratification
of dormancy takes place during the final stages
of seed development and thus the investigation of
at 4°C,” and that “seeds exhibit an improved
PGRs levels during this period should be dormancy and embryos excised from
revealing. . . .
The relatively few tree seeds in which PGR
research has been reported indicate the necessity
for increasing the number of species which are
selected for study. The lack of comparative work
between gymnosperm genera is especially noted.
However, this does not eliminate the need for
intensive, in-depth studies of individual spe- cies.
The detailed studies of Acer, Coryllus, and
Fraxinus provide examples of the desirability of
cooperative work on the same genera. (Leadem
1987, p. 85)

Kermode (1995, p. 274) asked, “What are the


im- portant factors or regulatory “cues” that
maintain embryos in a developmental state
and prevent them from undergoing a
premature transition to germina- tive events?”
Kermode then gave a lengthy discus- sion of
the role of abscisic acid (ABA): “Following
the termination of dormancy, seed
germination is completed, usually under
conditions different from those that
(originally) triggered the release from
dormancy.” According to Bewley (1997),
Much more needs to be learned about the key
processes involved in germination and dormancy.
Both germinat- ing and dormant seeds must
undergo many cellular and metabolic changes in
common after imbibition, and yet only the
embryos of the former emerge from their sur-
rounding structures. The real block to
germination in dormant seeds may occur at the
very last stage: radicle cell wall extension. Even
so, there may still be many steps that must be
completed between the perception of the signal
for dormancy breaking and the final emergence
of the radicle. In the past decade, most research
on the cel- lular aspects of dormancy has focused
on the secondary events, the metabolism of seeds
during and after release from dormancy, but to
little avail. New approaches that can be or are
being tried to an attempt to identify germination-
and dormancy-associated genes include T-DNA
mutagenesis, differential display, subtractive
cDNA hybridization, and the use of
nondestructive reporter gene technology. Perhaps
it is time to focus also on the primary events:
perception and transduction of the dormancy-
breaking signal. Finally, we need to determine
how radicle extension occurs, the ultimate
manifestation of germination. (Bewley 1997, p.
1063)

Taylor et al. (1993, p. 120) determined


that cold treatment affected gene expression,
168 Douglas-fir: The Genus Pseudotsuga
dormant seeds are capable of germination” (pp. to germinate. In another paper, Jarvis et al. (1997b)
255, 256). They also found that the proteins again supported the above.
encoded by the genes are hydrolytic and may have
a role in cold- induced dormancy breakage.
Forward et al. (2001) found that metalloproteinase
and serine proteinase activity increased during
early stratification and that “degradation of major
soluble proteins probably oc- curs through the
action of multiple proteinases acting in a specific
development cascade” (pp. 626, 628). They noted
that during conifer seed development, large
amounts of lipids, carbohydrates and storage
proteins accumulate (p. 625). Misra (1994) found
that lipids made up about 50% and 30% of
Douglas-fir megagametophyte and embryo dry
weight, respec- tively, and that proteins made up
about 12% and 10% of Douglas-fir
megagametophyte and embryo dry weight,
respectively (p. 360).
Ching and Ching (1973) suggested that “ger-
mination specific MRIVA is essential in breaking
dormancy” (p. 3). They also showed that
stratifica- tion resulted in an increased energy
change and adenosine triphosphate (ATP) content
of Douglas-fir seed, while Malavasi et al. (1986,
p. 35) showed that stratification raised ATP 13x in
the embryo and 6x in the gametophyte. Sorensen
(1971, pp. 10, 12) noted that seeds with “white
mutant” embryos did not require stratification and
germinated more rapidly than normal seeds.
Jarvis et al. (1996) reported that late
embryogen- esis abundant (LEA) protein genes
are expressed during dormancy breakage of
Douglas-fir seeds and suggested that the proteins
may be involved in moisture relations. They
stated that their report was the first paper
presenting evidence for these genes in
gymnosperms and that the level of genes is
enhanced with stratification and correlated with
depth of dormancy (p. 565). They also noted that
Douglas-fir seed dormancy is seated in the
structures surrounding the embryo, which does
not have en- dogenous dormancy. In a second
paper, Jarvis et al. (1997a) suggested the above
and noted that “while de novo synthesis of ABA
was important for the expression of dormancy,
four weeks of chilling only led to a 36% reduction
in endogenous ABA” (p. 457). They also noted
that methyl jasmonate stimulated dormant seeds
Chatthai and Misra (1998) discussed seed erences to Douglas-fir seeds are included,
storage protein genes that occur in the late however, especially regarding their seed proteins.
embryo genesis of Douglas-fir seed. Although She noted the three broad classes of seed proteins:
they diverge from similar proteins in “structural proteins, associated with membranes
angiosperms, the translational products share and ribosomes,
a structured homology and indicate a common
ancestor for angiosperm and gymnosperm 2S
storage protein genes. Misra and coworkers
dis- cussed the biochemistry of proteins in the
zygotic embryo of Douglas-fir (Owens et al.
1993) during the early germination of
Douglas-fir seed (Misra 1993) and during
embryo genesis dormancy release and
germination (Misra 1994).
Owens et al. (1993) reported the following:
The ultrastructural, histochemical, and
biochemical development of the post-fertilization
megagametophyte and the zygotic embryo of
Douglas-fir megagametophyte and embryo
development were studied from fertiliza- tion
until seed maturity, a period of about 71 days.
The most important morphogenetic events
occurred during the first 43 days. During this
time lipid bodies and pro- tein bodies increased
rapidly in the megagametophyte. Lipids,
proteins, and starch became evident in the em-
bryo toward the end of the morphogenetic phase.
The subsequent embryo maturation phase
showed slight increases in size and number of
megagametophyte lipid bodies and protein
bodies, as well as an increase in pro- tein body
complexity. Later, in the mature seed, lipids and
proteins were distributed uniformly in the
megaga- metophyte. Starch was abundant in
some regions of the embryo but not abundant in
the megagametophyte. In mature seeds soluble
sugars made up 2 and 3%, proteins 16 and 11%,
and lipids 60 and 45% of the megagameto- phyte
and embryo dry weight. They continue (p. 823)
from 14 to 43 days after fertilization “dry weight
and soluble sugars did not increase but lipid and
especially protein increased in the combined
megagametophyte and embryo.” Owens et al.
(1993, p. 816)

Misra (1993) found that “mobilization of


storage proteins is rapid between 4 and 6 days
of germina- tion” (p. 77). The preceding has
discussed a number of aspects of “dormancy.”
However, the most ap- propriate summary is
a statement from Taylor et al. (1993):
“Despite many years of research on tree seeds
the mechanisms underpinning dormancy
break at low temperatures are virtually
unknown” (p. 120). In an excellent, detailed
review, Misra (1994) discussed several
aspects of seed biochemistry, but much of the
material is about angiosperms. Some ref-
Chapter 7. Seeds 169
but as the dormancy was reduced, gibberellin lev- els
(2) enzymes e.g., those required for mobilization fell; stratification had little effect on cytokinins.
of storage reserves, and (3) storage proteins which Applications of either GA4 or GA7 did not stimulate
are utilized during seed germination and seedling
growth, thus supplying the necessary free amino
acids and amino nitrogen” (p. 362). With
reference to conifers, she included the following:
In the conifers examined so far (including Douglas-
fir) the protein bodies consist of globoid and
crystalloid inclusions embedded in a buffer-soluble,
amorphous proteinaceous matrix. In mature seeds of
Douglas-fir (Pseudotsuga menziesii) crystalloid
aggregates of highly in- soluble polypeptides make up
70-80% of the total storage proteins and are located in
protein bodies. Douglas-fir also contains a 55-kDa
complex as major storage protein. (Misra 1993, p.
364)
In discussing the LEA proteins, which we
previously
noted, Misra observed,
Generally the LEA proteins are hydrophilic and
contain a large number of uncharged as well as
hydroxylated amino acids arranged in conserved
protein domains. They are believed to stabilize
other proteins and pos- sibly membranes, thus
protecting seed tissue during desication and
dormancy or cellular disruption upon subsequent
rehydration. Most of the LEA genes can also be
induced in other plant parts by exogenous ABA
application (rab genes—ABA responsive genes) in
the absence of water stress. Other stress treatments,
such as wounding, salt, and cold, can elicit
expression of these genes. Therefore, the function
of these proteins may reflect a common protective
role in plant cells when stressed. (Misra 1993, p.
366)
The gene expression studies in conifers have
focused mainly on storage proteins and recently
on LEA proteins. Misra (1993) noted that
“maximum expres- sion of crystalloid protein
genes in conifer mega- gametophytes was
achieved during the embryonal mass stage before
the formation of the meristogentic regions of the
embryo” (p. 368).

Growth regulators
Bradbeer (1988) discussed seed dormancy and
ger- mination in some detail, but the material is
confined to examples of angiosperms. He
suggested that the chilling that causes the
breaking of dormancy in ha- zel is associated with
gibberellin synthesis. Taylorson and Hendricks
(1977, p. 337) noted that “gibberellins are active
stimulants of seed germination.” While Taylor and
Wareing (1979a) showed that, at first,
stratification increased gibberellins in Douglas-fir,
This is because our present understanding of the
germination of unstratified seeds, possibly control of germination is unsatisfactory. Our knowledge
of dormancy regulation— by PGRs or other means—in
because it did not penetrate the seed coats. In
tree seeds is superficial and fragmented. Seed
addition to the gibberellins, growth regulators dormancy studies tend to be scattered among a limited
commonly include the auxins, cytokinins, and number of species, and only rarely is
abscisic acid (the latter is frequently classed
as a growth inhibitor).
Gibberellins
In her review of the role of plant growth
regulation in seeds, Leadem (1987, p. 2) noted
the following:
Gibberellins actively stimulate seed germination
of many species of angiosperms and
gymnosperms. Over 60 different GAs have been
identified, but those most frequently used
exogenously in forest tree seeds are GA 3, GA4,
and GA7. Dormancies with chilling and light
requirements are often overcome by GAs, and
synergistic interactions between phytochrome
and applied GA have been cited. Increases in
endogenous GA are usually, although not always,
coincident with the termination of dormancy in
seeds undergoing dormancy-releasing treatments
such as stratification and light. Although ger-
mination of seeds is frequently associated with
increased GA levels, it is unclear at what stage of
the germination process this increase occurs. It
can only be stated that exogenous GA relieves
dormancy and is an endogenous component of
many tree seeds—a priori evidence for a
significant role for GA at some point in the
termination of dormancy, or the initiation of
germination. (Leadem 1987, p. 2)

Only two of the references she cited mention


Douglas-fir: the preceding paper by Taylor
and Wareing (1979a) and Richardson’s (1959)
contribu- tion, which reported that 5 ppm GA
accelerated germination of non-stratified
Douglas-fir seed incu- bated in light at 20°C.
Although the review (1987) cited no papers
that related any of the remaining growth
regulators to Douglas-fir seed dormancy,
Leadem (1987) commented that “in depth
studies of plant growth regulators (PGR’s) in
conifer seeds have not been reported and it is
not legitimate to extrapolate results with
angiospermous seeds to those of
gymnosperms” (p. 83), concluding,
The germination of tree seeds appears to be
controlled by a variety of external and internal
factors. PGRs fig- ure prominently among these
factors, but the range of mechanisms by which
PGRs control is mediated may vary considerably
—from physical to metabolic. Although significant
support exists for PGR involvement in the
regulation of angiosperm seed dormancy, the
evidence for gymnosperms is still inconclusive.
170 Douglas-fir: The Genus Pseudotsuga
any one species investigated extensively. to germinate, concluding that the longer the chilling
Coordination between various studies is difficult
because in some case investigations have been treatment the lower the capacity of ABA biosynthesis
confined to entire seeds, while in others, only to parts and/or the higher is the catabolism of ABA in seed
of seeds. Adding to the confusion is the extrapolation transferred to 15°C (p. 317). They also
from studies of exogenous PGRs to experiments in
which endogenous PGR levels are cor- related with
dormancy-breaking events. Because of the uncertainty
of PGR permeability, or the extent of their
metabolism, it is impossible to deduce from
exogenous PGR experiments the effective
concentrations of PGRs at their site of action.”
(Leadem 1987, p. 84)

Abscisic acid (ABA)


Bianco et al. (1997) showed that seed dormancy
was correlated with ABA levels; levels of this
compound fell as dormancy increased, and seed
covering tissues allowed denovo-synthesized
ABA to accumulate in the embryo and maintain
its dormancy. Corbineau et al. (2002) noted that
the content of ABA in the embryo or
megagametaphyte tissue fell with dor- mancy
release, further catabolism of ABA by the seed
increased, and synthesis decreased with dormancy
release. Additionally, “the inhibitor effect of ABA
on seed germination was more effective at 15°C, a
sub- optimal temperature than at 25°C thermal
option” (Corbineau et al. 2002, p. 318).
According to Bewley and Black (1994),
“maturation drying is the normal terminal event in
the development of many seeds, after which they
pass into a metabolically quiescent state” (p. 122).
They suggested that such desiccation results in a
decline of ABA content, possibly lowering the
embryo’s sensitivity to ABA, and that ABA may
contribute to desiccation tolerance at the end of
the maturation phase. These data are from
experiments with angiosperms; we are aware of
no similar work that has been done with Douglas-
fir.
Corbineau et al. (2002) cited several reports
that
supported the involvement of ABA in seed
dormancy and others that disputed such a
relationship. They argued that the “involvement of
ABA in the ger- mination of gymnosperms is
poorly documented” (p. 314). Corbineau et al.
stated that “dormancy in Douglas-fir is seated
mainly in the embryo; however, the seed coat and
megagametophyte may also play a role” (p. 316).
They showed that chilling greatly lowered the
level of ABA and increased the ability of the seed
found that breaking dormancy was associated maturation period in cells of both storage tissue and
embryo
with decreased sensitivity to ABA.
At maturity, dry seeds have many conserved but
Germination inactivated systems. Some of the systems, such as sol-

Bewley and Black (1994) defined germination


as beginning “with water uptake by the seed
(imbibi- tion) and end[ing] with the start of
elongation by the embryonic axis, usually the
radicle” (p. 1). As defined by Leadem (1996),
germination is the “re- sumption of active
growth in the embryo, which results in the
protrusion of the embryo from the seed and
the development of the embryo into an
independent plant” (p. 18). Kolotelo (1997)
wrote that “seed germination is recognized by
the emergence of the radicle from the seed …
when the cotyledons emerge, germination is
termed episeal and this is characteristic of the
conifers. Much of the research concerned with
Douglas-fir seeds has investigated internal
and external changes associated with ger-
mination” (p. 25). Koller (1972) wrote that
the ger- minating seed has few requirements
that must be satisfied by the environment:
This leaves the immediate post-germinative
growth ac- tivity with virtually few and simple
requirements from the environment. One of these
is an adequate moisture supply, yet which would
not interfere with the gaseous exchange which is
essential for aerobic respiration and adequate
supply of metabolic energy. Another such
requirement is for “normal” temperature, i.e.,
within the range which is suitable for normal
growth of the more mature seedling. (Koller
1972, p. 3)

In contrast, studies of the germination of


coniferous seeds have been dominated by an
as yet unsuccessful attempt to generate a
parameter that will estimate the capacity of a
given seed lot to produce numbers of vigorous
seedlings. Examples of these efforts include
the work of Campbell and Sorensen (1979),
and Thomson and El-Kassaby (1993).
Ching (1973a, p. 76) discussed seeds in
general (her work has included angiosperm
seeds and seeds of Northwest conifers,
including Douglas-fir):
The germination process in turn can be divided
into three distinct, yet overlaping and
interacting, phases:
(A) reactivation of preexisting systems, (B)
synthesis of enzymes and organelles for
catabolism of reserves, and
(C) synthesis of new cellular components.

Reactivation of conserved systems from the


Chapter 7. Seeds 171
radicle then emerges through the seed coverings, using substrates
uble enzymes, transfer ribonucleic acids (tRNAs), and transported from the storage organs. Again, this phase is
mitochondria area, easily reactivated by water at opti- conditioned by early reactivation processes and catabolic
mum temperature and proper atmospheric conditions. activities in the storage organs and is controlled by environmental
Ribosomes, however, need to be dissociated to conditions. (TM Ching 1973, p. 80)
subunits or reactivated to become functional in
protein synthesis. Long-lived messenger ribonucleic
acids (mRNAs) are generally not available as
templates in dry seeds until they are freed from
protein or exposed during imbibition. The in vivo
sequence of activation of these systems can be
illustrated in the wheat embryo. The reactivation
process often requires an energy supply from the
biological fuel, adenosine triphosphate (ATP) which
is usually low in dry seeds as shown by lettuce seeds.
The ATP content is soon built up during imbibition
through glycolysis, fatty acid oxidation, respiration,
and synthesis of adenosine diphosphate (ADP). This
reactivation phase appears to be accomplished during
imbibition or soon after. The major events of this
phase probably are:
• First, reactivating conserved systems to start
basal cellular metabolism, such as turn-over type
main- tenance, protein synthesis, glycolysis, fatty
acid oxidation, solute and ion transport,
cytoplasmic streaming, respiration, and so on;
• Second, building up ATP content for various syn-
thetic activities after imbibition; and
• Third, providing enough substrates for respiration
and protein synthesis. Sugars and fatty acids are
common substrates for respiration. The sugars are
conserved and easily produced from starch by the
action of pre-existing phosphorylase and b-
amylase. The fatty acids, usually pre-existing, are
the hydro- lytic product of triglyceride by
conserved lipases.

This reactivation phase sets the metabolic wheel in


mo- tion. The faster the initial speed and the more
functional the conserved systems, the greater the
germination force.
Synthesis and sustenance of enzymes and
organelles
for catabolic degradation of reserves
This phase occurs mostly in the storage organs of
the seed, usually in response to an instruction (for
example, hormone or long-lived mRNA) either pre-
existing or coming from the embryo or embryonic
axis. The activ- ity or quantity of the protein-
synthesizing machinery (ribosomes, mRNA’s and
tRNA’s – polysomes), mito- chondria, glyoxysomes,
enzymes, coenzymes, cofactors, substrates, and other
factors increase during and after imbibition, reach a
peak at about 50% exhaustion of reserves, and then
decline to complete exhaustion of reserve. (Ching
1973, p. 76)
Synthesis for anabolism in embryonic or embryo axis
The pre-existing substrates and biochemical
systems
and the rapidly increased respiration during
imbibition
provide ATP for protein synthesis, which in turn
supplies enzymatic, structural, and soluble proteins
required for the myriad processes of growth. The
During germination of Douglas-fir seeds, TM
Imbibition of mature dormant seeds Ching (1963a) found that “a marked reduction of
initiates several metabolic processes that saponifiable lipids observed in germinating seed
result in a trans- formation of storage after stratification was accompanied by an
products into energy and anabolic increase in sugars and starch. With germination,
substrates for germination and seedling nitrogen content per individual seed remained
growth. In a series of papers, Te May constant, while fresh weight increased 3 to 4 fold
Ching and colleagues described some of and dry weight decreased 5% to 15%. Some
these reactions that occur in Douglas-fir decrease in oligosac- charides and free fatty
(TM Ching 1959, 1961, 1963a, 1963b, acids, and a slight increase
1965, 1966, 1968; TM Ching and Fang
1963;
TM Ching and Schoolcraft 1968). TM
Ching (1959) reported that the normal
course of germination has four stages:
The 1st, imbibition, is accomplished in 12
hours with almost a linear increase of
respiration and water uptake. The 2nd
stage, to 36 hours, is characterized by a
constant respiratory rate and R.Q., and a
temporary cessation for further water
uptake. The 3rd stage prior to emergence
consists of a gradual increase of water
uptake and re- spiratory rate, and a rapid
rise of R.Q. to approximately
1.15 at the time of radicle emergence
(seedling stage A). The 4th stage consists of
a further increase of respira- tory and water
uptake but with a decline of R.Q. in the
seedling and the attached partially digested
endosperm. The 1st stage is common to
most seeds studied; the uptake of water
probably creates the proper intracellular
environment or a degree of hydration which
is essential for enzyme systems to function,
and in the meantime the raised respiration
probably supplies the energy
requirement for the onset of the germination process.
The 2nd stage has been observed by
other workers at the stage prior to seed
coat rupture (24). The low diffusion
coefficient of an O2 (15) for a uniform rate
of oxidation throughout the tissue, the low
concentration of essential substrates and
cofactors for enzymatic reactions in
catabolic as well as anabolic work, and/or
the slow rate of enzymatic activities
might be the limiting factors. It may be
considered as an antephase of
mobilization.
The 3rd stage is the active
mobilization of energy source and cellular
components preparatory to the later stages
of germination that are accompanied by
true growth in terms of cell number, cell
size, and tissue differentiation.
The 4th stage after radicle emergence
symbolizes a remobilization for cotyledon
emergence, as evidenced by a lowering of
R.Q. and increased respiration and water
uptake.
During the above, fatty subtances are
oxidized with the production of ATP.
(TM Ching 1959, p. 560)
172 Douglas-fir: The Genus Pseudotsuga
in reducing sugars and total extractable lipids was and neutral lipase, fourfold during germination”
found in stratified seed.” (TM Ching1968, p. 482).
TM Ching (1963b) also reported the following: Table 7.1 shows the changes found in Douglas-
In contrast to angiosperm seeds, little is known fir cone scales and seed during their development
regard- ing the metabolic pattern of formation, (TM Ching and Fang 1963, p. 551). Cones were
quiescence, and germination in gymnosperm seed.
collected during the summer from a single tree
The major food
reserve in Douglas-fir seed was found to be fats and near Corvallis, Oregon (Table 7.1). They found
an increase of carbohydrates accompanying a rapid that at maturity—156 days after pollination—the
decrease of glycerides was observed during
seed had “12 mg in dry weight, 40% fatty
germination of the seed. (TM Ching 1963b, p. 722)
substances, 30% nitrogenous com- pound, 20%
Total fats decreased rapidly with germination from fiber, 4% other carbohydrates including a trace of
36% to 12% of the dry weight, which also decreased
from starch and 4% minerals” (p. 553). Other data
13.1 mg to 12.1 mg per individual seed. Glycerides demonstrated that “seed absorbed 4 times as much
were utilized during germination and a diminution of glucose as the scale in early developmental stages,
86% to 59% of the total fats was found. Acetone-
insoluble phos- pholipids increased gradually in the then declined to twofold at the last stage of the
early stages, then rapidly at later stages of experiment. Synthesis of fat and cellular compo-
germination. They increased from 5% to 25% of total nents from labeled glucose was demonstrated, and
fats, of which fatty acids com- prised approximately
30% at any stage of germination. A small reduction the rate of fat synthesis in the seed increased with
from 2.5% to 1.5% of total fats, then an increase of maturity. A preferential utilization of carbon-1 for
5% was observed in the fraction of free fatty acids respiration and carbon-6 for synthesis was
during germination. (TM Ching 1963b, p. 727)
indicated by the differential radioactivity in
Earlier, Ching (1961) had found that Douglas- various fractions” (p. 554). We have noted that the
fir seeds contain 40% to 45% fatty substances, of germinating seed has few requirements that must
these 70% glycerides, 7% acetone precipitable be satisfied by the environment; nonetheless,
phospho- lipids, 3% free fatty acids, and 1% several environmental factors do influence
nonsaponifiable material. In two papers, she germination, including light, moisture, and
discussed the composi- tional changes in Douglas- temperature.
fir seed during germina- tion (TM Ching 1965,
Light
1966), reporting that “lipids, proteins and reserve
phosphorus compounds in these gametophyte In the section discussing flowering, many plant
were utilized for the synthesis of carbohydrates, re- sponses are governed by “phytochrome,” a
structural components, and soluble compounds in chemical whose form is affected by the duration
the seedling”(1966, p. 1313). and quality of light. The germination of seeds
In the final paper of the series, Ching noted affected by phy- tochrome (and light) is termed
that the highest activity of lipases “was found to “positively photo blastic.” The following reports
be with the heavy fat bodies.” She also found that present evidence for this condition in Douglas-fir
Douglas- fir seed contained “35% lipids, 32% seeds.
protein, 29% fibers (seed coat), 1.8% minerals, Johnson and Irgens-Møller (1964) found that
1.7% starch and sugars, 0.2% RNA, 0.04% “the rate of germination of unstratified seed of
nucleotides and 0.03% DNA.… Acid lipase Pseudotsuga menziesii (Mirb.) Franco was increased
activity increased sevenfold

Table 7.1 Change of weight, moisture, and nitrogen content of developing Douglas-fir cone scale and seed.

Fresh weight (mg) Dry weight (mg) H2O content (%) N content (% of dry weight)
Days after pollination Scale Seed Scale Seed Scale Seed Scale Seed
75 710.5 25.4 137.6 3.9 80.8 84.6 1.38 3.01

85 662.9 26.8 173.0 5.7 73.9 78.2 0.98 2.99


95 578.4 22.4 174.0 6.6 69.7 70.8 0.86 3.28
105 711.1 25.1 236.8 11.9 66.7 59.4 0.91 4.53
Figures listed are the average of two determinations which deviated less than 10% of the average (Ching and Fang 1963).
Chapter 7. Seeds 173
are reversibly photochromatic, red/far-red light
significantly by interruption of the dark period absorbing lipoproteins” (p. 160). Bewley and Black
with two hours of white light compared with the (1994) noted that “almost all light-requiring seeds
rate obtained without interruption of the dark
period but with the same total amount of light per
24-hour cycle” (p. 200). These results were
limited to trials in which the temperature was
below 25°C and con- firmed earlier studies by
Richardson (1959). Johnson and Irgens-Møller
also tested the effects of red and far-red light on
the germination of Douglas-fir seeds. The data
showed that red light stimulated, whereas far-red
light inhibited germination at 23°C, but that
raising the temperature to 30°C did not eliminate
the retarding effect of far-red. These data agree
with the findings of Pons (1983), and suggest that
where plant competition creates shade and shifts
the red/ far red in the light to far red, seeds
respond to this signal by failing to germinate. The
foregoing agrees with early work by Allen
(1941c), which indicated that light treatment just
before germination was stimulating to unstratified
Douglas-fir seed. The stimulating effect of light
was greater at higher tem- peratures, but because
neither temperature nor light were controlled,
drawing definitive conclusions is difficult.
Similarly, workers at the California Forest
Experiment Station found that photoperiods (even
with intensities less than 100 foot-candles) greatly
stimulated the germination rate of unstratified
seed, but not of stratified seed (1957, p. 19).
Recognizing these effects, the standard
germination protocol for seed germination tests
include 9 hours daily of fluorescent (red-light
rich) light.
Interestingly, an experiment designed to
deter-
mine whether seed orientation affected germina-
tion showed that the greater light transmission of
one side of the seed was not responsible for more
rapid germination (Sorensen and Campbell 1981).
Other reports examined light stimulation of pine
and spruce seed (Ackerman and Ferrer 1965,
Taylor and Wareing 1979b). Devlin et al. (1995)
discussed a range of phytochrome species,
describing red-far reversibil- ity as “the classical
hallmark of phytochrome action;
. . . responses of this type are considered to reflect
the so-called low fluence response (LFR) mode of
phytochrome action” They continue, “phytochromes
many of the seeds will germinate near the freezing
have coat-imposed dormancy” (p. 236). point and the radicles continue to grow at that
Although the Service Testing Manual (Stein tempera- ture (0–2°C).” Allen (1962a) found that
1966) recommended a daily exposure of 8 moisture at 60% to 70% on a dry-weight basis
hours to cool, white fluores- cent light was adequate
(possibly to inhibit moulds), Baskin and
Baskin (1998, p. 412-414) noted that Douglas-
fir seed germinate equally well in light or
dark. Bewley and Black (1994) gave a lengthy
discussion of possible phytochrome activity in
seeds but, unfortunately, no conifers were
mentioned. Li et al. (1994) noted that daily
photo periods increased both total ger-
mination and germination speed of
nonstratified Douglas-fir seed, while only the
rate of germination increased for stratified
seed. Alosi et al. (1990, 1992) found that,
unlike for angiosperms, light is not nec-
essary to promote chlorophyll and other
pigments in Douglas-fir seed.

Moisture
Leadem (1988) noted that an adequate
moisture supply is necessary for germination;
Bewley (1997) argued that “by definition,
germination incorpo- rates those events that
commence with the uptake of water by the
quiescent dry seed and terminate with the
elongation of the embryonic axis” (p. 1055).
According to Koller and Hadas (1982),
The first requirement for germination is water for
hy- drolysis, as a medium for translocation by
diffusion, for hydration of enzymes, cell
membranes and organelles to their conformation,
and finally to provide the driving force for cell
expansion that is initiated by germina- tion.
Consequently, the amount of water required
for germination itself is that which suffices to
bring the relevant tissues of the seed to the
adequate level of hydration. These are minute
amounts of water. (Koller and Hadas 1982, pp.
402-403)

Most research on the role of water in the


germi- nation of Douglas-fir seed has been
devoted to the moisture level of seeds during
the stratification or pregermination phase.
Early on, moistening the seed to an optimum
level was considered sufficient, but this
procedure may allow radicle emergence
before the release of dormancy of many seeds
in a given seed lot. Such actively elongating
seeds cannot be used in projects to produce
seedlings. Allen (1967,
p. 63) noted that “after extended stratification
174 Douglas-fir: The Genus Pseudotsuga
for the stratification of Douglas-fir seeds, and that affect the moisture content of the embryo (5% of
seed stratified with a moisture content (MC) of whole seed weight) or of the gametophyte; drying the
70% retained the effects of stratification after a seeds to MC of 25%
period of drying. Seed stratified at 40% moisture
germinated less well after drying. Most complete
germination at 15°C occurred after 70% moisture
during stratifica- tion. The more recent papers are
primarily concerned with the moisture level of the
seeds and with drying either during or after
stratification. Copeland (1976,
p. 160) noted that presoaking, followed by drying,
has been shown to benefit many agricultural
seeds. Danielson and Tanaka (1978) stratified
Douglas-fir seed and then dried it to three
moisture levels (no drying, air drying, and oven
drying) before storing at different temperatures
for 3, 6, and 9 months. Seeds at highest
moisture content did not store as well as those at
the lowest. The stratification effect was not lost
after the 2°C storage. They found that "air drying
immediately following stratification (to 37%) is a
method which may be used to safely store seeds
germinating during extended stratifica- tion
periods” (Danielson and Tanaka 1978, p. 16).
Edwards (1986) reviewed several studies over a
range of species (including Douglas-fir) where
dry- ing after stratification, followed by cold
storage of the dried seed, increased seed
germination parameters, particularly germination
speed. Edwards believed this latter effect was a
result of more synchronous germination in seed
lots because the drying step “does not prevent
those processes that accompany dormancy
removal from occurring” (p. 164), so that the
more dormant seed have time to catch up dur- ing
the storage phase of the procedure, and, when
moisture stress is relieved by a free water supply
in a favorable germination environment, all seeds
that
can germinate do so synchronously.
In two papers, Malavasi et al. (1985, 1986) dis-
cussed the effects of stratifying Douglas-fir seed
for 28 days (fully saturated, 3°C) followed by
storage at 3°C at a moisture content of 25%, 35%,
or 45% for 1 or 3 months. Stratified seeds were
dissected after storage and the effects of moisture
treatments mea- sured for the seed coat, embryo,
and gametophyte tissue. These latter data showed
that drying seeds to 35% moisture content did not
reduced the embryo MC from 50% to 32%. They during the treatment. Paulsen (1996) found that
reported the following: extending the stratification of Douglas- fir seed at
Three months of storage did not alter moisture
MCs of about 30% (fully imbibed MC is
distribu- tion within seeds. Stratification reduced
the germination percentage of interior-source
seeds but hastened germi- nation speed for seeds
from both coastal and interior sources. Redrying
stratified seeds to 35 and 25 percent moisture
content increased seed vigor and seedling length
and dry weight remarkably, a response similar to
the <<invigorating effect>> reported to improve
seed performance for other types of plants.
Storing stratified seeds, without redrying them,
for 1 or 3 months gener- ally reduced seed vigor,
as reflected by germination speed and seedling
length and dry weight, yet redried seeds stored
no better than nondried. Levels of bio- chemical
compounds studied were strongly correlated with
germination speed. Results suggest that it would
be advantageous to redry seeds to a range of 25
to 35% moisture content directly before sowing
to produce vigorous seedlings or allow
expression of stratification benefits. (Malavasi et
al. 1985, p. 371)

Malavasi et al. (1986) reported the following:


Certain biochemical attributes (adenosine
phosphates, nucleic acids and total nucleotides)
were analyzed in Douglas-fir [Pseudotsuga
menziesii (Mirb.) Franco] seeds and seedlings
from a coastal and an interior seed source in
Oregon to explore how seed stratification,
redrying and storage interact to produce vigorous
seedlings. Seeds were stratified at 45% moisture
content (MC) and then redried (to 35 or 25%
moisture content) and/or stored (for 1 or 3
months) in a range of treatment combinations.
Stratification increased ATP 13-fold in the
embryo and 6-fold in the gametophyte; energy
charge rose from 0.4 to 0.8, and RNA increased
60 to 80% in the embryo and 150 to 300% in the
gametophyte. Redrying stratified seeds to 35 or
25% moisture content increased RNA and DNA
greatly in the embryo but not in the gametophyte.
Storing redried seeds generally reduced all
biochemical attributes. Stratified, redried seeds
produced the most vigorous seedlings, though
their biochemical attributes showed no constant
advantage, possibly due to their rapid
metabolism. However, the benefit of
stratification and redrying was not preserved in
stored seeds of either source. (Malavasi et al.
1986, p. 35)

Responses of the seeds were similar, whether


or not the embryos were dried. This similarity
suggests that control of seed metabolism is
not seated in the embryo.
Jones and Gosling found that stratifying
seed at a target MC of about 2% below
saturation resulted in the same response as the
regular stratified procedure, but that the seeds
remained surface dry and did not germinate
Chapter 7. Seeds 175
diminution may be the reason Douglas-fir seed is
about 40%) yielded good germination data. Her viable under natural conditions for only 1 year after
data also demonstrated that maintaining moisture seed fall.
at predetermined levels was critical. In his review,
Jensen (1996) noted that the critical moisture
content for controlled moisture stratification was
commonly between 3% and 8% below full
hydration (p. 297) and for Douglas-fir, 32% to
36% (p. 302). He also noted that if MC was kept
in this range, stratification may be prolonged
significantly (p. 299).
Moller et al. (1999) found that prechilling
Douglas- fir at 32% MC for 34 weeks at 3°C
enhanced germi- nability and germination speed;
subsequent storage at 6.7% MC and 3°C for 6
months did not have a detrimental effect on
germination, but raising MC to 8.1% did. These
data are interesting, particularly when compared
to those of Malavasi et al. (1985), who found that
differences in seed moisture were not necessarily
reflected by differences in embryo moisture:
“comparisons between seeds prechilled at
controlled MC before storage and those strati- fied
with the traditional method resulted in better
performance of the first ones, both in the
laboratory and the nursery.”
Data from Gosling et al. (2003) showing differ-
ences in individual seed moisture content at target
moisture are also interesting in this respect. In
their study, Douglas-fir seeds were moist-
prechilled at target MC of 10%, 15%, 20%, 25%,
30%, 35%, and 40%
for 2, 4, 8, 16, 32, 48, 64, or 120 weeks and then
incu- bated at 15°C (the temperature that most
strongly reflected dormancy release). The
procedure yielded several relationships: (1) optimal
dormancy breakage occurred in Douglas-fir seeds
at moisture contents on a fresh-weight basis of
between 30% and 35% and, after prechills of
durations between 25 and 48 weeks, the MC of
most of the combinations tested were between
15% and 30%. (2) The authors made the
interesting observation that dormancy breakage of
Douglas-fir seed under dry, warm storage
occurred only at MC below 20%, but that the
maximum effects of stratification occurred at MC
>25%. (3) Finally, they noted that prolonged
stratification at higher MC (30%-40%) resulted in
seed death, possibly because of respiratory
diminution of carbohydrate reserves. This
The temperature range within which germinability
A summary of the above studies suggests is maximal is considerably wider than that which is
conducive to maximal rates. The optimal
that, for Douglas-fir seed, a moist prechill
temperature is the range within which both
where the moisture content of seed is held parameters are maxi-
near to, but below saturation, is most effective
in releasing dormancy without premature
germination. These and other studies,
however, fail to describe the basic mecha-
nisms of dormancy.

Temperature
Laboratory manuals containing protocol for
Douglas-fir seed germination recommend
temper- atures between 20°C and 30°C and a
pretreatment (stratification) at 4 oC. As noted
previously, there is an interaction between
light and temperature, which produces varying
germination. Danielson and Tanaka (1978)
reported that stratified Douglas-fir seed
germinated more rapidly at 5°C than at 2°C.
Allen (1960a, 1962a,b) reported that the
length of the stratification period had a strong
effect on seed germination in the subsequent
incubation period— the longer stratification
permitted germination at a lower incubation
temperature. He also noted a strong difference
in the temperature requirements for coastal and
interior Douglas-fir seed (Allen 1960a,
1962a,b). The coastal seed required higher
tempera- ture. Sorensen (1991) reported
similar results: seed germinated at an
incubation temperature of 15°C required 84
days of stratification for full germina- tion,
whereas seed incubated at 25°C required only
21 days of prechill.
According to Gosling (1988), "the effect of tem-
perature on the germination of seeds of many
species has been frequently examined.
However, a statement by Lang (1965) that
‘precise and unequivocal infor- mation on the
complete germination temperature range of
different species is surprisingly rare’ is
almost as true today as it was 20 years ago”
(Gosling 1988, p. 90). His data showed that
non-stratified seeds did not germinate at
10°C or 35°C and that the maximum
germination increased with tempera- ture
between those points. Stratification
increased germination between 15°C and
30°C; no stratified seed germinated at 35°C.
Koller and Hadas (1982) argued as follows:
176 Douglas-fir: The Genus Pseudotsuga
mal. Above this optimum, the progressive decline in hours at 20°C produced the most rapid and greatest
germination rate, and eventually also in germinability,
total germination.
probably involve a time-dependent thermal
denaturation of proteins and a phase change of the cell Besides the absolute effect of temperature upon
membranes. (Koller and Hadas 1982, p. 404) seed germination noted above, several workers
The effect of low temperatures was reported have reported an interaction of temperature with
by Hawkins et al. (2003). They found that other environmental factors or with previous seed
freezing dry Douglas-fir seed to −19°C had little history; Allen (1941c), Richardson (1959), and
effect on seedling germination (p. 1237). Freezing Johnson and Irgens-Møller (1964) all noted a
tolerance was much lower for imbibed seed, light-temperature interaction at temperatures
however, and reached a minimum at the stage of below 25°C. We further speculate that seeds
germination when hypocotyls were rapidly falling from the cones in the spring may require
elongating. Cold hardiness of seed lots and high temperature to germination, since they were
germination rate were negatively correlated. not exposed to natural cold moist conditions on
Studies have been conducted to define three the soil surface.
major effects of temperature upon Douglas-fir While Finnis (1950) reported that stratified
seed: (1) effect of temperature; (2) effect of high seed germinated at lower temperatures than did
tempera- ture upon viability of seed; and (3) effect similar unstratified seeds, Allen and Bientjes
of a range of lower temperatures during storage (1954) noted that the optimum germination
upon the subsequent viability of the seed. This temperature of seed stratified for 1 week was
last effect of temperature will be discussed in the higher than that of seed stratified for 6 weeks. The
section on seed storage. latter is probably an ex- ample of Vegis's (1964)
Early field observations (British Columbia theory that as an organ is released from dormancy
Forest Service 1940) indicated that the effect of it is able to resume growth under an increasingly
soil tem- perature on seed germination was far wide range of temperatures. Although the limited
greater than the effect of light. Similarly, data describing the direct effect of environmental
Lavender (1958) noted that seed placed on a factors such as light, moisture, and temperature
south-facing slope germinated more rapidly than does permit defining an optimum range for each
seed on nearby easterly, westerly, or northerly for seed germination, much more work is required
slopes. if we are to understand the interaction of
The first published account of a study designed environmental factors in stimulating or inhibiting
to quantify the temperature necessary for seed growth.
optimum germination is that of Allen (1941c), Only three references were found (Willis 1917,
who suggested 20°C as a standard for germination Hofmann 1917, Morris 1936) which dealt with
tests, but pre- sented no data demonstrating the the effects of extreme heat upon Douglas-fir seed
effect of higher or low temperatures upon viabil- ity. Willis (1917) discussed a study
germination rate in total. Slightly later work by conducted with relatively crude equipment, but
the same author (Allen 1947) demonstrated a which yielded data generally confirmed by later
greater germination rate and total for seed lots work. He concluded that the possible damaging
held under a 38°C to 15°C thermoperiod than effect of temperatures employed in drying
obtained under thermoperiods of either 27°C to cones was a function of the cone moisture.
15°C or 21°C to 12°C. However, the experimen- Thus, 49°C for 10 hours greatly reduced the
tal control of the temperatures was rather crude viability of seeds from cones that were green
and no mention was made of the duration of any when the drying started, but had little ill effect
of the temperatures employed, so it is difficult to on similar seeds extracted from cones that were
compare these data with other data in the par- tially dry before the heat was applied.
literature. Later workers (Allen and Bientjes 1954, Hofmann’s (1917) concern was with the effects
Holmes and Buszewicz 1958, Johnson and Irgens- of forest fires upon the subsequent germinative
Møller 1964, Jensen and Noll 1959) agreed that capacity of seed in the duff. His laboratory trials
either about 25°C constant or a daily cycle of 8 demonstrated that at 71°C in dry heat, the
hours at 30°C and 16 seedcoat began to darken and pitch began to
exude from the seed. Seed in moist
Chapter 7. Seeds 177
The viability of seeds (frequently noted in this work)
heat did not show the shriveling of endosperm or is based on the results of one or more germination
the marked darkening that seed might if exposed
to temperatures as high as 82°C for 10 hours in a
dry atmosphere without losing viability;
temperatures above 60°C for 10 hours severely
reduced the germi- native capacity of the seed
which was exposed in a very moist atmosphere.
Field tests with seeds placed about 2.5 cm under
forest floor materials showed that a hot slash fire
did not damage the seed; nor were the
temperatures of the seeds raised above 49°C.
In contrast to the above, Morris (1936)
utilizing more elaborate experimental apparatus,
which may have measured the actual seed
temperature more accurately, found that
temperature as low as 55°C for 3 hours would
reduce the viability of seeds with an initial MC of
77% (dry seed) while temperatures of 45°C were
damaging to seed with an initial MC of 60% (wet
seed). Morris’s tests were conducted with seed in
closed glass tubes, while Hofmann apparently
employed open lots.
Allen (1958a) employed several kiln tempera-
tures, from 40°C to 60°C, to dry preconed cones.
He found that kiln temperatures between 40°C
and 52°C produced seeds with significantly higher
(>90% vs. about 70%) germination capacities, but
that the cones dried at 55°C or 60°C yielded seeds
that germinated more rapidly than did those from
cones dried at the lower temperatures (50%
germination = 12 to 25 days for stratified seed).
Finnis (1955) employed laboratory conditions to
test the effects of 37 days of 10°C, 16°C, and
24°C temperatures upon total germination after a
subsequent 24°C for 32 days; his data
demonstrated that 10°C limited and 16°C partially
limited germination.
The above confirms the hypothesis that one
cause of low germination of spring-sown seed
during one year was due at least partially to
temperatures in June. With projected global
warming, soil tempera- tures may not be low
enough to satisfy stratification requirements and
would hence impact field germi- nation
(Lavender, unpublished data).

Seed Tests
Germination tests
(15.5-35°C) as the most favorable temperature for
tests, cutting, or biochemical (“quick”) vigor. In Douglas-fir seed. (Jensen and Noll 1959, p. 107)
1816, the first standard germination test was Jensen and Noll (1959) also reported erratic responses
developed in Europe. Today, the testing of to prechilling and a negative effect from presoaking.
Douglas-fir and other seeds follows the
specifications in publica- tions of regional
(Stein 1966) and international seed testing
bodies, such as the Association of Official
Seed Analysts and the International Seed
Testing Association (1985). Reported
germination trials with Douglas-fir seed
(Toumey and Stevens 1928) were in soil on
greenhouse benches. They noted that “blue”
Douglas-fir seeds germinated quickly and
completely, but that “green” Douglas-fir seeds
from California were much slower. Holmes
(1951, p. 10) discussed a variety of tests in
Europe, many of which were abandoned, and
reported a close correlation between
germination tests and tetra-zollium tests via
biochemical trials for several slowly
germinat- ing species. Jensen and Noll (1959)
discussed early testing of Douglas-fir at the
Oregon Cooperative Seed Laboratory, noting
the following examples (see (Tables 7.2, 7.3,
7.4):
Testing Before 1950
In the search for a method for testing Douglas-fir
seed, several factors were considered. The
temperature of germination chambers; the need
for and the length of prechilling, and the use of
presoaking as a treat- ment before germination.
Before 1950, fewer samples of Douglas-fir seed
were received for testing than in later years
[Table 7.3].
Temperature was checked by using three
alterna- tions (20 and 30°C, 10 and 30°C, and 15
and 30°C) and one constant temperature (20°C).
In the alternations, the lower temperature was
held for 16 hours daily and the higher for 8 hours
daily. Fluorescent light of 75 to 100 foot-candles
was supplied at the higher temperature. A
mechanical counter planted the seeds in closed
transpar- ent containers (plastic or glass) on a
moist substratum (usually sand, otherwise paper
toweling). Two to four replicates were tested
under each condition. Average germination
results for the 13 lots tested at these tem-
peratures were 68, 70, 56, and 47%, respectively.
The constant temperature (20°C) was not
satisfactory; all lots except three gave incomplete
germination at this temperature. At the
alternating temperatures, tests of only three
samples had ungerminated seeds remaining
which appeared to be sound when dissected at
the end of the test period of about 42 days. The
first two sets of alternating temperatures gave
better results than did the 15-30°C alternating
and the 20°C constant temperatures. The Woody-
Plants Seed Manual (5)2/ recommends 60- 95°F
178 Douglas-fir: The Genus Pseudotsuga
They recorded the effects of prechilling by years, did not exceed 8% for any lot. Slightly higher results
were obtained by prechilling. For example, in January
for a total of 840 samples (Tables 7.2, 7.3, and the average germination was 63.7% with no prechill
7.4). and 66.9% with prechill.
The second class was characterized by a simple
A somewhat similar pattern was shown by the 114
dor- mancy, as evidenced by a definite response to
sam- ples of Douglas-fir seed received in 1951-52.
prechilling and the sound seeds in replicates not
Although the biological history of individual samples
prechilled. Most samples received in August,
varied, the trend toward periodic dormancy is
September, October, and March followed this pattern.
interesting. The survey by months showed that
Prechilling increased
response to treatment fell into several classes. The
germination by 4 to 80%, with an average gain of 38.9%.
first comprised samples that showed slight response to
The third class was characterized by complex dor-
prechilling. Samples tested in July, November,
mancy, as shown by the sound seeds at the end of the
January, April, and May showed little difference
test period under both testing methods. This dormancy
between methods. No sound seeds remained at the
end of the tests. The difference between methods

Table 7.2 Average percentage of germination without prechilling and with various periods (from Jensen and Noll 1959).

Prechilling time in weeks at 5°C


No prechill 2 3 4 5
Year Number of samples Germination (%)
1941 9 52.6 60.6 — 64.1 61.3
1944 6 65.3 — 67.7 — —
1945 10 59.8 67.9 — 70.7 —

Table 7.3 Average germination by years using no chilling or prechilling of Douglas-fir seed (from Jensen and Noll 1959).

Year Number of samples No chilling (%) Prechilling % gain from prechilling


1941-42 10 52.6 60.6 8.0
1943-44 8 65.3 67.7 2.4
1945-46 31 59.8 67.9 8.1
1947-48 6 61.5 57.8 -3.5
1948-49 14 61.5 57.0 -4.5
1949-50 45 67.8 74.1 6.3
1950-51 161 64.7 70.6 5.9
1951-52 114 58.2 70.7 12.5
1952-53 22 74.5 76.6 2.1
1953-54 69 69.9 72.5 2.6
1954-55 237 74.9 77.6 2.7
1955-56 192 76.1 78.5 2.4

Table 7.4 Comparison of germination results by range and difference between no treatment and for 2 weeks of prechill.

Number of samples
Higher, with no prechill Same Higher, with prechill (%)
(%)
Year Total samples 1–7 8–15 16+ 1–7 8–15 16+

1949–50 45 6 3 3 — 20 5 8
1950–51 161 33 2 3 3 60 46 14
1951–52 114 21 5 — 2 40 21 25
1952–53 22 6 — — 1 13 2 —
1953–54 69 9 4 — — 38 17 1
1954–55 237 46 17 2 16 108 42 6
1955–56 192 36 5 — 14 100 36 1
Totals 840 157 36 8 36 415 169 55
Chapter 7. Seeds 179
tests should be limited to seed that has received this
probably was different from those encountered in
other months because neither method gave
pretreatment. They urged revision of ISTA standards.
complete germi- nation. Such samples were
received during February and early March. No
difference in germination was found between the
two methods, but all tests contained sound
ungerminated seeds at the end of the testing period.
Further study may develop a method that
gives the optimum germination of the exceptional
sample that shared a type of dormancy that does not
respond completely to present methods.
Results of tests during September and October
illus- trated another problem in testing coniferous tree
seed. The total of percent germination plus percent
sound seed obtained under unfavorable test conditions
generally does not equal the percent germination
obtained under more favorable conditions. Without
prechilling, the
average germination plus sound seed was 57.6% and
with prechill was 72.4% for these 16 samples. (Jensen
and Noll 1959, p. 108)

Evidently, some seeds decayed that would have


ger- minated under favorable conditions, which
argues for testing in the most favorable
environment. The results so obtained did not
necessarily correspond to the percentage of plants
found after field sow- ings, however; Jensen and
Noll (1959) reported that a 2-week prechill period
was adequate for dormant seed germination and
that no one treatment gave good results for all
seed lots. They also recommended that tests
should include two, 100-seed samples with no
pretreatment and two similar samples given a 2-
week prechill.
There have been a number of studies for
evaluat- ing the potential of seed to produce
viable seedlings, given a favorable environment
(Anon 1954; Jensen and Noll 1959; Campbell and
Sorensen 1979; Stein 1966, 1967; Ching and
Jensen 1959; Thomson and
El-Kassaby 1993). Ching and Jensen (1959, p. 52)
ex- amined the variation in seed weight, purity,
moisture content, and viability of a Douglas-fir
seed lot. Their results showed that these seed
parameters varied among storage containers and
by how many seeds were in a container. They
recommended sampling each container for each
lot. Gosling and Peace (1990,
p. 796) noted that laboratory germination and
prob- ably field germination of Douglas-fir seed
benefited from a pretreatment of 2 weeks of
stratification, and that this condition is sufficiently
common to warrant that laboratory germination
however. One interesting variant of the standard
Thomson and El-Kassaby (1993, p. 125) germination test, and one that may increase the
noted that, for Douglas-fir seed, the variations in the germination rate, was
probability of a seed germinating on a given
day is a better measure of the speed of
germination than either the time to 50% ger-
mination or the main daily germination of the
most vigorous component of a seed lot. This
conclusion, however, was based on the
performance of a seed given a relatively short
(2 wk) stratification period and leaves open
the question as to whether it would be the
same after longer stratification. The authors
did show that a fast-germinating seed lot with
low viability could have the same peak value
(P.V.), a parameter suggested previously as a
similar value combining germination speed
and total viability, as a seed lot with slow
germination and high viability (p. 130), and
hence would be of questionable value for
container nurseries. The thinning in containers
in nurseries is generally guided by the speed
of germination and therefore may result in
seedlings with different genetic makeup.
Working with loblolly pine seed (which,
like Douglas-fir, can be dormant), Barnett and
McLemore (1984, p. 161) found that total
laboratory germina- tion was a better estimate
of nursery performance than was peak value
for all but slowly germinating seed lots. In
contrast, Czabator (1962) argued that, “as a
rule, nursery germination generally closely
approximates the laboratory figure only in the
case of rapidly germinating seeds of high
quality” (p. 387). He cited data from several
workers—as well as his own—showing that
germination speed and total germination were
superior estimates of nursery performance
than was total viability alone (p. 388).

Standard germination test


Stein (1965, p. 22) reviewed the value of the
standard viability test, which includes a
germination trial and the determination of
the percentage of pure seed and seed
weight. Laboratory and field germina- tion
results correlated well: the points were on or
close to the 45° line. Obviously, total
germination of Douglas-fir seed in the
germination chamber provides only a weak
approximation of nursery performance,
180 Douglas-fir: The Genus Pseudotsuga
germinative vigor.
reported by Sorensen and Campbell (1981). They
found that the speed of germination was affected
by the side of the seed that was in contact with the
germination medium. They reported the
following:

The physical basis for the polymorphism resides in


the shape and development of a Douglas-fir
(Pseudotsuga menziesii (Mirb.) Franco) seed. A
wingless, mature seed is flattened and ovate,
approximately 8 mm x 3 mm x 2mm. The seed coat is
differentiated into three layers, of which the outer is
continuous with the adaxial surface of the ovuliferous
scale. Separation of the seed wing from the scale
results from dissolution of the middle lamella beneath
the ovule and seed wing. When the seed is separated
from the scale, the side previously adhering to the
scale is lighter in colour and somewhat flatter than the
upper side. Because the seed is somewhat flattened, it
tends to rest on a more or less even surface with the
scale side up or down. (p. 467)
The evidence for an adaptive basis for the
variation pattern comes from the observation that
both mean germination rate (day-1) for a provenance
and the dif- ference between germination rate scale-
side-up and scale-side-down increased with
distances from the ocean. In addition, the
differences between germina- tion rates scale-side-
up and-down were very closely related to mean
germination rate for provenance. In other words,
provenances whose seeds had the most rapid
germination rates also showed the greatest response
to orientation, and both germination rate and
response to orientation increased with increasing
distance from the ocean. Length of frost-free season
and mean an- nual precipitation decreased regularly
with increasing distance from the ocean in the area
sampled. (Sorensen and Campbell 1981, p. 470)

Cutting test
Perhaps the simplest seed-evaluation procedure is
the cutting test: the seed is dissected lengthwise
and the contents examined. This technique can
determine if the embryo is fully formed; the
technique is par- ticularly useful before cone
harvest in determining the seed maturity. The
embryos should be at least 90% of the length of
the embryo cavity and should be examined for the
degree of insect damage and to determine whether
the megagametophyte tissue is mature (it should
be firm and white; Eremko et al. 1989, pp. 25–26).
Finally, the seeds are exam- ined several hours
after cutting. No shrinkage of the
megagametophyte tissue should be present,
indicating tissue maturity (which may occur after
embryo maturity). This tissue condition is
important to maximize seed storability and
Biochemical quick tests a trained and skilled analyst.

Because the standard germination test for


Douglas- fir requires several weeks to
complete, a significant effort has been made
to develop "quick tests" or a more rapid
scheme of evaluating seed viability.
Hydrogen peroxide
Ching and Parker (1958) reviewed
previous tri- als, and presented their test
results of Douglas-fir seed viability with this
chemical. In their trials, seeds with the radicle
end excised were soaked for 5 days in a 1%
solution of hydrogen peroxide (H2O2)
maintained at daily alternating temperature of
20°C for 16 hours and 30°C for 8 hours. At
the end of this time, seeds with radicle
elongated 1–8 mm were tallied as viable. The
average viability determined by this method
was 3.3% higher than that shown by the
standard germination test, tetrazolium
chloride (2,3,5-triphenyltetrazolium chloride).
Flemion and Poole (1948) and Copeland
and McDonald (2001) reviewed work with
this chemical. Viable seeds were identified by
the reduction of tet- razolium chloride, a
colorless chemical, to bright red formazan by
the action of dehydrogenase enzymes.
Healthy, viable seeds were identified by red
staining. This test requires only a few hours,
but the technician must be skilled at
interpreting the results. Flemion and Poole
(1948, p. 252) noted that although there was
general agreement between seed viability, as
measured by tetrazolium, and that determined
by the excised embryo test, the differences
that existed were enough to question the
accuracy of any single tetrazolium test.
Buszewicz and Holmes (1952) reviewed
the lit- erature, particularly the German trials,
and their laboratory results, which showed a
correlation of 0.931 between 119 trials of
tetrazolium chloride and parallel germination
tests of Douglas-fir seeds. They found the
tetrazolium procedure to be superior to a test,
using several different chemicals (p. 142), and
concluded that “the regression equations will
assist in obtaining a more reliable estimate of
seed quality than has been possibly hitherto”
(p. 150), providing the tests are conducted by
Chapter 7. Seeds 181

Excised embryo
may not do well under sub- or supra-optimal
Flemion (1948) presented data for 21 families conditions. Seedlots that germinate well, and produce
and 87 species (but not Douglas-fir) showing a seedlings under a wide range of conditions, are said to
be of high vigor. Thus, vigor tests try to predict seed
cor- relation of 0.949 between viability, as performance under a variety of conditions. No single
measured by excised embryos, and parallel tests has been developed to quantify the vigor
germination tests (p. 235). Flemion (1938, 1941) attribute, especially for tree seeds, but the tests that
are used are based on other attributes that distinguish
reported a fair correlation between the results of more-vigorous and less- vigorous seeds. Vigorous
excised embryo tests and germination tests for seeds germinate rapidly and compete better for water,
light and nutrients over a wide range of environmental
Douglas-fir seed. According to Copeland and
conditions, especially tempera- ture. While respiration
McDonald (2001, pp. 136–137), this test is varies according to the internal moisture level, and the
particularly useful for woody shrubs or trees stage of germination, higher respiration may signal
impaired physiological activity,
whose dormant seeds may require germination
i.e. lower vigor. In some vigor tests, seeds are
tests extending for months. They further noted incubated at low temperature (the cold test), or high
that although many seed laboratories routinely temperature combined with high humidity
(accelerated aging tech- nique), to compare vigor
conduct these trials, one disadvantage is that these levels. Not only do vigorous seeds germinate better,
tests re- quire very skilled analysts. but they are more resistant to disease organisms.
Vigor tests are often used to predict field germination,
Seed-vigor tests since field conditions are less certain and controllable
than in the laboratory. Comparisons between
McDonald (1993, p. 93) observed that seed vigor laboratory tests and field results for the same seedlot
test- ing began with developing the standard continue to bedevil seed users. (Edwards et al.,
unpublished manuscript, pp. 56-57)
germination tests and reflected seed performance
under ideal conditions. Although such testing may Ching (1973a) discussed seed vigor for all seeds:
not estimate seed performance under adverse field Seed vigor may be defined as a potential for rapid and
conditions, it has become a routine method for uniform germination and fast seedling growth under
testing field performance capability. No general field conditions. Germination processes can be
divided into three distinct, yet overlapping and inter-
references were made to vigor testing of Douglas- acting stages: reactivation of conserved systems from
fir seed in Copeland and McDonald (2001), which maturation period; synthesis for catabolic activities,
includes an extensive review of this subject. mainly in storage organs; and, synthesis for anabolic
ac- tivities in the embryo. Growth involves three
In an unpublished manuscript on the collection sequential components, namely increase in cell size,
and handling of Pacific Northwest forest tree cell number and degree of differentiation. While the
basic pattern of germination and growth is
seeds, Edwards et al. discussed germination and
programmed by the ge- netic makeup of the species in
seed vigor as follows: question, the eventual expression of the pattern,
however, often is modified by environmental
Germination refers to the reactivation of physiological conditions under which seeds are grown, harvested,
processes that result in the growth of the embryo to processed, stored, treated and planted. Therefore, the
form an independent seedling. “Reactivation” is the status of vigor in a seed lot stems from the interaction
operative word for almost all PNW species because a of all parameters involved. Some factors assert more
majority of the seeds entered the state of greatly- stress on seed vigor than others, depend- ing mostly
reduced metabolism known as dormancy as they upon the degree and the timing of stresses. Not all the
matured, and more so when they were stored. When sequential biochemical events of seed germination and
mature seeds naturally detach from the mother tree in embryo growth are precisely identified yet. We may,
the fall, or are collected by foresters, their moisture however, summarize some experimental data in the
level is quite low, an evolutionary step designed to literature and from our own laboratory in-
limit immediate germination with winter approaching.
dicative of some specific factors affecting seed vigor.
Whereas a small proportion of the naturally-falling
Of the conserved biochemical systems in seeds,
seeds may germinate if temperatures are favorable,
en- zymes, proteins, mitochondria, ribosomes and
most seeds lie dormant in the forest floor over winter.
mem- branes appear to be major sites of aging caused
(p. 53)
by poor storage conditions. High humidity and
All forest tree seeds vary widely in their ability to
temperature in seed storage also reduce substrates for
produce vigorous, healthy seedlings. Whereas
early enzyme activities. Seed maturity seems to affect
germina- tion tests determine seed lot quality as the
the functional status of long-lived messenger
potential of a sample of individual seeds to develop
ribonucleic acids (mRNA) which encode germination
into independent seedlings under ideal germination
events. Over-drying of seeds denatures proteins and
conditions, some lots
inactivates enzymes. Low ger- mination temperature
fails to reactivate the pre-existing
182 Douglas-fir: The Genus Pseudotsuga
ribosomes and mRNA, thus eliminating polysome for- the use of non-stratified seed; thus, it is normally
mation and protein synthesis. Any reduction, lesion or
defect mentioned above will lower seed vigor as any used in tests of Douglas-fir seed (Lavender
physiological process results from a cooperative effort 1978a). Although agriculturists have recognized
of component biochemical systems. The weakest link the posi- tive effects of stratification on seed
or the lowest substrate, effector, coenzyme or co-
factor concentration often limits enzyme activities, germination since the middle of the 17th century
germination and growth. (Ching 1973a, p. 73) (Evelyn 1664), the last 50 years have seen several
Ching (1973a) argued that it "is difficult to define papers adding to various aspects of stratification
seed vigor to suit everyone. But to a practical to the pattern of germination in Douglas-fir seeds.
agri- culturist or a seed technologist, vigorous Allen and Bientjes (1954) found that the details of
seeds have the potential to germinate rapidly and the stratification that yielded the highest seed
uniformly after planting, and the emerging germination varied with seed lot. They
seedlings have the ability to grow vigorously recommended 6 weeks at 0° to 2°C, with MC of
under general, sometimes relatively adverse, field about 60%. In a series of reports, Allen (1960a,
conditions” (p. 76). She con- cluded, “the 1962a,b) examined the effects of the variation of
common correlation of seed respiration and seed factors involved in stratification on the germina-
vigor indicates that poor growth must be related to tion parameters of treated seed. In his 1960 report,
some impairment in mitochondrial activ- ity” (p. Allen noted that the rate of germination increased
84). In another report, Ching (1973b) noted that with the length of the stratification period and that
“ATP content in imbibed seeds is significantly incubation temperatures showed less difference in
correlated with seedling size in fatty, starchy and germination with longer stratification. Allen
proteinaceous seeds, and it indicates viability in (1962a) reported the following:
seed lots. ATP content thus appears to be a useful An initial moisture content of 60-70% was confirmed
biochemical index of seed vigor” (p. 400). She ar- as an effective one for stratifying more and less
dormant seed of Douglas fir. Furthermore, the moist
gued that “the level of ATP in plant tissue appears seed can be dried, after stratification, at room
to be a very sensitive index to environmental and temperature for 24 hours and stored at 0-2°C in closed
developmental changes” (p. 402). Data describing containers for extended periods without loss of
germinative capacity or decrease in rate of
ATP levels may possibly predict seedling vigor. germination. The initially drier seed is sensitive to
much further drying particularly the more dormant
Stratification lots.
For these reasons an initial moisture content of
Stratification treatment, whether in the laboratory about 70% is recommended. Previous unpublished
or under field conditions, involves exposing moist data have indicated that seed so treated can be
seeds to temperatures of about 2°C. The net effect partially dried after stratification for 20–150 days and
then stored for as long as one year without loss.
of such treatment is to relieve the state of (Allen 1962a, p. 307)
dormancy and prepare the seed for uniform,
vigorous germination. A study by Lavender According to Allen (1962b),
(1958a) demonstrated that field stratification had The results suggest that good quality seed can be
dried after stratification, if not needed, and stored at
a significant effect on subsequent seed 0-2°C for at least one year with little loss. It can be
germination, and it served as a guide to aerial stratified for 120 days or longer without loss of
seeding of the Tillamook Burn. A large fire in the viability in most cases and then either used,
continued in cold storage with gradual drying, or
Oregon Coast range was seeded after February 1. dried down to 10% quickly and re-stored at 0-2°C.
No germinates were found the following year and Even limited room-temperature storage at 10%
the area was subsequently planted. These results moisture is satisfactory for period of a few months.
(Allen 1962b, p. 490)
were confirmed by Carmichael (1957). Warming
trends predicted for the coming decades may Edwards (1986) reported four main
result in soil temperatures, particularly at low prerequisites for successful chilling:
elevations, that are too warm to properly chill First a source of moisture is required to rehydrate seed
Douglas-fir seed and, hence, result in lower tissues, which in conifer seeds have usually been
dried to moisture contents below 10% for preservation
natural regeneration. in cold storage. Uptake of water allows essential
Stratification is used in lab tests because it re- biochemical changes to begin. Second, temperatures
between 1°C and 5°C are required to favor certain
sults in more rapid germination than occurs with
biochemical changes and morphological
developments, while delaying embryo
Chapter 7. Seeds 183
and thus result in more uniform post-storage
elongation (and sprouting) in individual seeds which
have lost their dormancy. Low temperatures also
germination. In a lengthy review, Hagery (1978)
reduce decay caused by microorganisms and prevent discussed a range results from hydration/dehydra-
dam- age by respiratory overheating. Third, good
aeration is necessary to prevent carbon dioxide
accumulation and minimize heat accumulation.
Fourth, treatment must be of the correct duration.
(Edwards 1986, p. 152)

He reviewed work demonstrating that the


germina- tion performance of the seeds of a
number of conifers including Douglas-fir may be
improved by subject- ing them to a period of cold
storage with reduced moisture after a “standard
stratification period.”
A study by El-Kassaby et al. (1992) using seed
collected from several trees demonstrated that a
stratification period as short as 3 weeks resulted
in dramatic differences in the germination param-
eters of the different seed lots. In a series of
studies, Sorensen (1980, 1991, 1996) found that
longer stratifi- cation periods produced higher
germination values, particularly when the
incubation temperature after stratification was
low. Sorensen noted that seeds stratified for 128
days produced higher germination when the
incubation temperature was 15°C, but not 25°C
(1980); and that seeds stratified for 84 days
germinated better at 15°C than those stratified for
21 or 42 days (1991). Finally, he reported that
seed germinated on cool, but not warm, nursery
soils (temperatures not given) germinated better
after than seed stratified for 14 or 33 days (1996).
Sorensen also reviewed data describing
stratification trials with a range of conifers,
including Douglas-fir, that reported the same
results (p. 197).
The previous discussion has focused on the
effect of low temperature storage on the
physiology of presumably fully inbibed seeds,
although Edwards (1986) noted that “little
attention has been given to controlling seed
moisture levels before, during, or after
prechilling” (p. 155). About three decades ago,
however, a series of trials were reported that
discussed the effects of drying seeds after a
period of prechilling or stratification before
returning the seeds to storage. The drying was
implemented to reduce the undesired sprouting of
seeds after lengthy moist prechilling and to
provide the time necessary to break the
dormancy of all the seeds in a given seed lot
followed by 12 weeks of stratification at a
tion treatment on subsequent seed controlled MC of about 8% below that of fully
germination, and noted that such treatment imbibed seed, allowed high and rapid germination
may allow embryo devel- opment, while of seed, particularly at an incubation temperature
preventing germination. According to of 15°C. She also showed that the seed could be
Edwards (1986), dried back to 8% MC and stored
The traditional view of breaking seed dormancy
at full inhibition has been challenged in both
conifer and broad- leafed species and, while the
new procedures (drying back seed after a period
of prechill but before storage) present difficulties
in application on a practical scale. They offer
increased flexibility to the plant producer in
terms of timing the initiation of prechilling and
in its prolongation to negotiate operational
problems caused by climatic conditions,
equipment failure, changes in planting schedules,
and so on. Some evidence shows that through
seed moisture regulation, dormancy can be more
thoroughly overcome and higher germination
capacities are possible. But the most significant
advan- tage related to increases in germination
speed brought about by synchronous
germination. . . .
Moisture management in tree seeds may not
be a universal solution in all species, but it
appears to have a broadly based relevance that
has been largely overlooked until the last decade.
(Edwards 1986, p. 168)

Gosling (1988) demonstrated that


prechilling increased Douglas-fir seed
germination at all incu- bation temperatures
between 10°C and 30°C, con- cluding, that
"relative dormancy can be suspected
whenever (1) seeds germinate over a narrow
range of conditions, and (2) seeds germinate
slowly under any onset of conditions," and
that "relative dormancy can be confirmed if
pretreatment either; (1) widens the range of
temperatures over which seeds can germinate;
(2) improves the maximum percentage at
some, but not all temperatures; and (3)
increases the rate of germination" (p. 95).
A later report by Jones and Gosling (1994)
showed that maximum germination and
germination speed occurred after pre-chilling
Douglas-fir seed for 36 weeks at 10°C and
MC of 38% (2% below full im- bibition).
None of the seed extended their radicles, but
all seed were apparently primed. Jenson
(1996) discussed the theory of prechilling at a
controlled moisture content in detail, noting
that the procedure allows dormancy release to
occur, but eliminates the ability of
nondormant seed to germinate. Paulson
(1996) found that 2 weeks fully imbibed,
184 Douglas-fir: The Genus Pseudotsuga
without losing germination capacity. According to vidual seed components, such as seed coat,
Muller et al. (1999), embryo, and megagametophyte. Further, no
Two main conclusions can be drawn from seedling measurements were made of the effects of seed
emergence tests. First, the benefits of long pretreat-
manipulations upon compounds known to be
ment on germination at suboptimal temperature in the
laboratory were also found in the nursery when involved in Douglas-fir seed germination,
similar, nonoptimal temperatures are likely to prevail. physiology, or on the vigor of resultant
Secondly, prechilled seeds stored at the lowest MC
performed better than all the others also in the
germinates. In a pair of reports, Malavasi et al.
seedbeds, especially when rehydration before sowing (1985, 1986) addressed these deficiencies. The
was applied. For the poor performance of traditionally data clearly demonstrated that drying seed to 45%
stratified seeds, one can assume that dormancy
removal was incomplete because of the shorter MC did not affect the moisture content of the
chilling duration. On the other hand, in common embryo or gametophyte, whereas drying to 25%
nursery practice, stratification periods range from 6 to did result in lowering the moisture content of all
12 weeks to avoid the risk of premature germination
(Jones and Gosling 1994). seed parts. But the post-drying behavior of all
The combination of prolonged chilling treatments seeds was very similar, which leaves open the
at controlled moisture content with dry storage of question as to the mode of action of the moisture
nondor- mant seeds represents an improvement in
Douglas-fir seed technology as it offers several content of the embryo. Downie et al. (1993)
advantages. Extended prechilling allows germination reported that polyethylene glycol (PEG) priming
at the low spring soil temperatures in tree nurseries did not improve the speed of germination of seed
and reduces the risk of failure in the field as
germination is rapid and more synchronous, because of a range of conifers, but an in- teresting variant
the variability of dormancy de- gree within a seedlot on stratification showed that mem- brane tube
is completely overcome. Drying and storage of
invigoration at 30% MC was equivalent to
nondormant seeds offers the possibility to delay the
sowing date in case of bad weather; at any time seeds stratification. Again, the mean moisture content
can be withdrawn from storage and sown of the seeds was critical, but no measurements
immediately, since they are ready to germinate.
Moreover, stored seeds are surface dry so they flow
were made of the MC percentage of seed
prop- erly through sowing machines. Results obtained components. Accordingly, research has shown
with these Douglas-fir seedlots are encouraging, but that seed moisture content is critical to
further experiments are needed as differences among
seedlots might be considerable. (Muller et al. 1999, germination, but it does not explain why. To
pp. 176–177) date, the authors have found no work with
Sorensen (1998) sought to determine whether coniferous seed similar to that reported by
"various lengths of dry storage could be Vertucci and Leopold (1987) for angiosperm seeds.
confounded with genetic or other treatment effects Perhaps this approach would clarify moisture
in compari- sons of germination response" (p. 97). effects. Gosling et al. (2003) reported that both
The study findings were as follows: stratifica- tion duration and seed moisture
content affected the breaking of dormancy of
Storage main effects and interaction with family were
significant for storage periods up to 3 to 4 months, Douglas-fir seeds. Their results showed that
and for storage longer than one year if storage stratification periods of 48 weeks at 4°C
temperature was above freezing. With storage at –
12°C, the storage effect was small and the maximum
stimulated germination at all sub- sequent
effect was a delay in mean germination time at 15°C incubation temperatures from 10°C to 35°C. The
of 3.6 hours compared to the germination time of optimum MC was 25% to 35%. Interestingly,
fresh seeds. Family main effects were much larger. In
general, it appears that any confounding effect of fully inbibed Douglas-fir seed at 40% MC did not
storage on germination rate of Douglas-fir seeds will germinate as well as seed at 35%. The exception-
be small compared to genetic differences and to many ally long stratification period, together with the
treatment effects. Where small storage effects could
be important, it is probably best to use either fresh long chilling requirement (14 weeks) of Douglas-
seeds or seeds stored more than 6 months. Long-term fir buds (McCreary et al. 1990) is in accord with
storage should be at subfreezing tempera- tures, the hypothesis suggested by Sorensen to the effect
probably −10°C to −20°C. (Sorensen 1998, p. 97)
that the stratification requirements and the chilling
In all of the foregoing reports, seed moisture
requirements of species are correlated.
was reported on a whole-seed basis; no attempt
The foregoing has discussed several aspects of
was made to determine the effects of moisture
stratification. Although this technique has been
control on indi-
used and studied for many years, the nature
of the re-
Chapter 7. Seeds 185

sponses it stimulates is not known. As Taylor et The first report exploring factors affecting cone
al. (1993) summarized, and seed production of Douglas-fir is that of
Despite many years of research on tree seeds the Willis and Hofmann (1915). These workers
mecha- nisms underpinning dormancy-break at low discussed the results of a study designed to survey
termper- atures are virtually unknown. A favored
the relation- ship of the following variables to the
hypothesis has been that growth regulators . . . play a
role, either individually or in concert. However other production of cones and seeds (p. 142):
significant physiological changes occur during
stratification includ- ing increased gluconeogenesis
1. Altitude, high or low
(e.g., LaCroix and Jaswal 1967, Davies and Pinfield 2. Locality, northern or southern
1980) and modifications to respiratory pathways (Pitel 3. Soil, good or poor
et al. 1989, and references therein). Responses to
stratification are therefore numer- ous and complex. 4. Age, young, medium, or old trees
Many may be the result of dormancy break rather than 5. Size, small, medium, or large trees
causal events. (p. 120). 6. Health, good or poor because of conkiness trees
7. Stand, open or dense
Cone and Seed Production
They concluded that the “yield of cones per tree is
In the section discussing the incidence of
highest with medium aged, rather large trees,
flowering in Douglas-fir, we noted that there were
which grow in open stands, in warm localities” (p.
many factors associated with the differentiation
146). It should be noted, however, that Willis and
and development of micro- and megasporophylls.
Hofmann (1915) recognized that there may be
Naturally, there can be no cones that were not
exceptions to the foregoing. Furthermore,
preceded by female “flow- ers.” However, the
although they presented some data, there is no
reverse is not true: a good crop of flowers does
evidence that the experimen- tal design permitted
not necessarily result in abundant cones;
the examination of any factor independent of
accordingly, there are still more variables that
possible interaction or that any of the differences
affect cone production.
would be considered “significant” by modern
Issac and Dimock (1958) summarized the inci-
statistical analysis.
dence of cone crops as follows:
Cone crops in natural Douglas-fir stands are neither Seed size and anatomy
uniform in character nor regular in occurrence—either
among stands or among trees within a stand. Not all Douglas-fir seed is a medium-sized coniferous
trees produce cones, and even during years of good seed; the seeds of pines and firs are significantly
cone production a rather low percentage of individual
trees may be heavy cone producers. Similarly, a crop
larger than those of hemlocks and spruces.
considered good throughout the region may show ex- According to the Forest Service Woody-Plant Seed
tensive local variation ranging from bumper crops to Manual, the aver- age number of Douglas-fir (var.
failures. Moreover, cone production of both stands
and individual trees is affected by a complex of menziesii) seeds per pound ranges from 30,000 to
biological factors, the combined effect of which is 35,000 for California seed sources to 49,000 from
only partially understood. Therefore, anything but the British Columbia seed sources (USDA Forest
broadest ap- plication of generalized observations
may be seriously misleading. . . . Service 1948). In general, seeds from southerly
The quantity of Douglas-fir cone crops is quite locations and exposures are heavier than those
unpre- dictable—even at short range. The current from northerly sites and slopes. Sziklai (1969)
flower crop can be used as a rough indication of
prospective cone production for the end of the presents data describing cone and seed collections
growing season. Likewise, the ratio of pistillate to from 1,335 trees between 42° and 53° in British
vegetative buds may be useful to predict cone crops a
Columbia. There was a distinct clinal trend
full year in advance (Allen 1941a). Finnis (1953)
found this method somewhat impractical, however, whereby seed length increased from north to
due to the difficulty of obtaining a representa- tive south. Robinson (1963) showed that both seed
bud count sample—particularly in older stands. If
flowers (or reproductive buds) are scarce, then the
length and width and wing length and width
cone crop must necessarily be light. An abundance of differed signifi- cantly over an area of three
flowers on the other hand is no positive assurance that degrees of latitude by six degrees of longitude.
the crop will be heavy. Many things that may decrease
or even destroy the crop can happen between bud Allen (1961) found that seeds from coastal British
setting and cone maturity. (Issac and Dimock 1958, p. Columbia had a longer, more pointed and
11) pinched-in seed tip than those from the
186 Douglas-fir: The Genus Pseudotsuga
interior; interior seeds were more triangular, with Further, all species in a stand do not produce
greater sheen. White et al. (1981) found that indi- seed in a synchronous manner. Isaac found that
vidual seed in southwestern Oregon averaged the cone crop was "usually heavier on poor sites,
0.015 g (30,000/lb). According to Edwards and such as gravel soils or wind swept ridges, than on
El-Kassaby (1996), “seed size varies within and good sites” (p. 15). This is in agreement with
among parents. These variations are caused by Garman (1951): “Analysis shows all the
both environmental (i.e. position within seed production factors (of seeds, cones) were higher
cone, height and aspect of cone in the crown) and per unit area of the poorer site” (p. 5), referring to
genetic factors” (p. 482). The anatomical structure site index 110 and 140. But these observations
of conifer seeds, as shown below, has a contradict the data of Willis and Hofmann (1915,
dominant maternal contribu- tion. Loopstra and p. 150), who found heavier seed crops on good
Adams (1989) reported extensive genetic sites. They reported 9,500 sound seeds per bushel
variation in Douglas-fir in southern Oregon: of cones from trees on shallow, gravelly soil, as
“seedlings from drier southern units had heavier opposed to 14,700 seeds from a bushel of cones
seeds and were smaller (shorter, fewer branches), from trees grown on a good soil.
earlier bud set, more subject to second flushing Earlier, we noted references that described in-
and less damaged by frost than were those from creased flower production as a result of nitrate
northern units” (p. 240), and “families from fer- tilization. Smith et al. (1968) reported that the
higher elevations having lighter seeds and effect of fertilizer was erratic, but that successive
producing seedlings with earlier bud set and less applica- tions of nitropills over a period of 4 years
frost damage, on average, increased cone production by 26%. They noted,
than families from lower elevations” (p. 241). however, that their data "shows conclusively that
Tables 7.5 and 7.6 demonstrate the great yearly seasonal varia- tions resulting from largely
variation in Douglas-fir cone crops in both the unknown causes and individual tree differences,
Pacific Northwest and Europe. Gashwiler (1969, probably under strong genetic control are most
p. 390) found that the annual seed fall for one important” (p. 8). In contrast, Orr-Ewing (1965, p.
clearcut varied from 400 to 168,800 over a 12- 281) reported a strong flower- ing response to
year period. He noted that the percentage of good nitrogen fertilization on 8-year-old Douglas-fir
seed was positively correlated with the size of the seedlings.
seed crop. Reukema (1982, p. 249) found that the Several workers have reported that seed
seed fall in a young Douglas-fir stand varied from produc- tion varies with tree age. Eis and
none to 3 mil- lion per year over a 29-year period, Craigdallie (1981) suggested that cone production
and that heavy seed crops occurred at 1- to 4-year begins at 20 to 25 years; Lavender and Zaerr
intervals. (Perhaps the only general statement that (1985, p. 8) observed that female strobili on trees
can be made is that heavy cone crops do not occur 5 to 8 years old are not rare in Christmas tree
on successive years.) Neustein (1940) noted that 4 plantations. Isaac (1943, p. 18) reported cones on
of 18 seed years were heavy in Great Britain. trees 12 to 16 years old; Willis and Hofmann
Interestingly, even during heavy cone-crop years, 1915, p. 158) found seed cones on 14-year-old
not all trees produce. As Isaac (1943) points out: trees; Winjum and Johnson (1968, p.
Even in years of heavy cone crops not all the trees 4) did not cite a minimum age, but found cones on
pro- duce. When there are cones on trees in virgin
13-year-old trees; likewise Finnis (1950, p. 122)
stands, there are usually cones on open-grown trees,
but the reverse is not always true. Even open-grown found female buds on 20-year-old trees.
trees fail to flower some years, and individual forest- Isaac and Dimock (1958), Hofmann (1924),
grown trees probably produce seed less than half the
time, i.e., some trees may rest even during good seed
and Garman (1951) all argued that mid-aged
years. A record of the variation in seed-producing Douglas-fir produces the greatest quantity of seed,
habits of individual trees is furnished by the record of and noted that even trees over 600 years old are
a Douglas-fir seed-tree plot during 1927, when there
was a good seed crop. In this instance 24% of the seed produc- ers. Garman (1951) found that “in
trees bore a heavy crop, 41% medium, 23% light, and stands 40, 100, and 275 years old, dominant firs
12% none. A variation always exists regardless of observed for three crops averaged for each crop
crop abundance. (Isaac 1943, p. 15)
95, 135, and 1,340 cones per tree
respectively” (p. 31). Hofmann (1924) stated
Chapter 7. Seeds 187

Table 7.5 General rating of Douglas-fir (Pseudotsuga


menziesii) cone crops in Washington and Oregon from Table 7.6 The following sequence of seed years noted in
1909 to 1954. The Netherlands 1931–67.

Date Failure Light Medium Heavy Year Crop


1909 — — X — 31 Bad
1910 X — — — 32 Bad
1911 — — — X 33 Moderate
1912 — X — —
34 Bad
1913 X — — —
35 Fairly good
1914 — — — X
1915 — X — — 36 Very bad
1916 X — — — 37 Very bad
1917 — X — — 38 Very bad
1918 — — — X 39 Very bad
1919 X — — — 40 Poor
1920 — X — — 41 Very bad
1921 — X — —
42 Fairly good
1922 X — — —
43 Very bad
1923 — — — X
44 Very bad
1924 X — — —
1925 — X — — 45 Very bad
1926 — X — — 46 Moderate
1927 — X — — 47 Very bad
1928 — X — — 48 Good
1929 — X — — 49 Very bad
1930 — — X — 50 Poor
1931 — X — —
51 Very bad
1932 — — X —
52 Very bad
1933 — — X —
1934 — — — X 53 Very bad
1935 — X — — 54 Very bad
1936 — — — X 55 Failure
1937 — X — — 56 Average
1938 — — X — 57 Very bad
1939 — — X — 58 Fairly good
1940 X — — —
59 Very bad
1941 — — — X
60 Average
1942 — X — —
61 Very bad
1943 — X — —
1944 — — X — 62 Very bad
1945 — — X — 63 Very bad
1946 — — X — 64 Moderate
1947 — X — — 65 Very bad
1948 X — — — 66/7 Failure/poor
1949 — — X —
After vanVredenburch (1969) and LaBastida (1969).
1950 — X — —
1951 — — X —
1952 — — X —
1953 — X — —
1954 — X — —
Source: Isaac and Dimock (1958, p. 13).
188 Douglas-fir: The Genus Pseudotsuga
that 15-year-old trees may be expected to produce genes of Douglas-fir are distributed over its
4,000 seeds; 100- to 200-year-old trees, 40,000 range. Studies of the genetics of this species
seeds; and 600-year-old trees, 7,000 seeds. have demon- strated that this spread of genes is
Shearer (1985), Garman (1951), and Isaac (1943) quite limited, and the following observations of
all noted that trees with large crowns produced seed flight certainly confirm this conclusion.
more cones than those with relatively narrow Perhaps the first obser- vation of Douglas-fir
crowns. Isaac (1943) observed “that open grown seed flight was recorded by Hofmann (1924). He
trees with large crowns have yielded as high as found that wind seldom carried seed farther than
648 liters of cones in one picking, but the average the tree height and concluded that rodent
forest-grown tree, which has a narrow crown, activity was a more effective means of dis-
yields about 54 liters during a good cone year” (p. tributing seed widely. His hypothesis has not
16). Garman (1951, pp. 12–13) found that been supported by other workers, however; for
dominant trees produced 5 to 10 times the number example, Isaac (1943) noted that “animal
of cones of intermediate trees. And Kozak et al. movement of seed is not important”(p. 22).
(1963) noted that large trees produce more cones. Sissons (1933) also wrote about wind
The report of the Research Branch of the distribution of coniferous seeds, but did not cite
British Columbia Forest Service for 1939 (British Hofmann's (1924) work:
Columbia Forest Service 1940) stated the Where reliance is placed upon natural regeneration,
following: ”Cone crops may fail because of (1) the seeding habits of the species involved are matters
of fundamental importance. For species, which
rainy weather at time of pollination” [in the depend on the wind for transportation of their seed
section on flowering, we noted that Silen and from the par- ent tree to the place where the young
Copes (1972) contradicted this conten- tion]; and tree is to grow, an understanding of the factors
governing wind distribution of seeds would appear to
”(2) Preceding large crops, and (3) effects of be essential. The majority of the most important
weather on bud differentiation” (p. 16). According coniferous timber tree species of the world bear seeds
to Reukema (1961), "thinned stands produce adapted to wind distribution, but a search of forestry
literature dating to 1925 fails to bring to light any
much more seed than unthinned stands in good material of definite value on this subject. Most of the
but not poor seed years," and "sound seed observations which have been made seem to have
been based on studies of existing regeneration,
percent varies with size of seed crop by year, but accompanied by estimates (or guesses) as to the
is not affected by thinning" (p. 3). In a 10-year probable source of seed supply. . . (Sissons 1933, p.
period there was one good, one moderate, and one 119)
light cone crop. Cone production in other years (1) Seeds with marginal wings, such as redwood
or birch, fall very rapidly for their weight. In these
was negligible. In contrast to other reports, Roy seeds the center of gravity corresponds closely to the
(1960) found that the highest percentage of sound center of surface area. (2) Seeds with short broad
seed (50%) occurred during a poor seed year. terminal wings, such as the firs, have a less rapid rate
of fall for their weight. In these seeds the center of
Working with young, open-grown Douglas-fir, surface area is some- what removed from the center of
Winjum and Johnson (1962) found the largest gravity, though not far from the wingward end of the
long seed. (3) Seeds with terminal wings much longer
cones in the most vigorous portion of the crown,
than their width, such as the pines and spruces, fall
which probably had the highest photosyn- thesis. least rapidly for their weight. In these seeds the center
Tappeiner (1969, p. 174) found that cones were of surface area is considerably removed from the
center of gravity. (Sissons 1933, p. 121)
strong sinks and that reduction in foliage length
and diameter growth were correlated with their According to Isaac and Dimock (1958),
presence. As we noted in the beginning of this "Douglas- fir seed, like that of most conifers and
section, seed and cone production is affected by many broadleaf trees, has a wing that whirls the
a multitude of poorly understood factors; hence it seed as it drops from the cone and thus retards its
is both extremely fall. Whirling seed falls at the rate of 46 to 76 m
variable and unpredictable. per minute" (p. 3). A num- ber of researchers have
investigated seed flight of Douglas-fir and the
Seed flight
various factors affecting it (Boe 1953; Dick 1955;
Natural movement (generally by wind) of pollen Frothingham 1909; Garman 1951;
and seed is the principal mechanism whereby the
Gashwiler 1969; Haig et al. 1941; Dick 1955; Isaac
1927, 1929, 1930, 1943, 1949; Isaac and Dimock 1960;
Lavender et al. 1956; Pickford 1929; Roy 1957, 1960;
Chapter 7. Seeds 189
California seed sources (Sweet 1965) and those in the
Sissons 1928, 1933). These studies used a variety Pacific Northwest. Rafn (1915) summarized 25 years
of techniques to evaluate the various factors of seed testing, finding the average weight of
involved in seed fall. The following conclusions
are a sum- mary of their results.
1. Most of the sound seed falls under the stand
or
within 91 m of the stand edge.
2. Most of the seed falls between September and
March.
3. Variations in the terrain and in wind speed
make it impossible to predict the distance seed
will travel.
4. Down drafts and updrafts, particularly those
often associated with a fire, have strong
effects on seed movement. Some seeds are
blown tens of kilometers by winds associated
with fires.
5. Sufficient seed for regeneration falls within 125
m
of timber edge.
6. The distance of seed fall is correlated with the
height of the release point.
7. The incidence of seedlings, but not total
number, is related to the distance from seed
source.

Seed size and germination


In an earlier paper, El-Kassaby et al. (1992) stated
that “Douglas-fir seed germination, especially
germina- tion speed, is under strong maternal
control; how- ever, no relationship between seed
size, expressed by 1000-seed weight on either
germination capacity or speed was observed” (p.
49). Finally, working with seed collected from 40
widely spaced trees, Lavender (1958, p. 8) showed
no relation between seed size and germination
speed. El-Kassaby et al. (1992) noted that,
“germination of conifer seeds is the result of
much complex metabolic activity involving three
distinct genomes: the diploid embryo, the haploid,
maternally-derived, nutritional megagametophyte;
and the diploid, maternal seed coat” (p. 51) as
below. Clearly, although the genetic inheritance
of the embryo derives from both the male and
female par- ents, seed size is a function of the
female parent only. It should not be surprising,
then, that the correlation between seed size and
seed physiology is not strong.2 Sweet (1965) and
Ching and Beyer (1960) noted that Douglas-fir
seed weight increased with elevation in both
colonization, and that “the conflict, then, is between benefits of
Douglas-fir var. menziesii samples to be 13.3 small seed size, which promotes dispersal, and thus seed survival,
and large seed size, which promotes seedling vigor, and thus
mg and var. glauca to be 11.3 mg. Bialobok and seedling survival” (p. 407). We know of no data that examines these
alterna- tives for Douglas-fir.
Mejnartowicz (1970), working with more than
100 collections in the Pacific Northwest,
reported that germination increased with
increased seed size.
In two papers (1958 a,b), Lavender
reported that although the heavier Douglas-fir
seeds generally produced heavier seedlings,
this effect lasted for only one year. This
finding may have reflected that the larger
megagametophyte represented a better
environment for the embryo. Sziklai (1969),
working with collections from 1,335 Douglas-
fir trees between 42° N and 53° N, reported
that seed length—but not width or wing size
—increased clinally from north to south. With
the exception of wing width, the di- mensions
of coastal seeds were greater than those of
interior seeds. Working with seeds collected
from 15 open-pollinated trees in three separate
years, Silen and Osterhaus (1979)
demonstrated significant differences in seed
weight by year from individual trees and
showed that, although seed size and weight
were strongly correlated, neither was related
to seedling growth after 10 years. They also
showed that grading seeds by weight in a
given seed lot would reduce its genetic base.
Sorensen and Miles (1978) reported that the
seed weight of Douglas-fir increased with
distance of the seed source from the Pacific
Ocean, presumably in response to demands of
an increasingly arid environment.
In a similar paper, Sorensen (1983)
presented data for the Coast Range in
southwestern Oregon showing that seeds
collected on dry slopes were larger and
germinated more rapidly than did those from
more moist microsites. Baker (1972) reviewed
seed-size statistics for thousands of plants in
California, from herbs through trees, and
concluded that increasing seed weight of tree
seeds is primarily correlated with increasing
drought, the major limiting envi- ronmental
factor.

2. Winn (1998) reviewed studies indicating that seed size can


affect germination percent, rate, and seedling size for
angiosperms, concluding that, despite “strong selection
favoring large seeds, no evolutionary response is possible
because genetic variation is lacking or ismasked by
environmental variation” (p. 1543). Schupp (1995) speculated
that seed size may not be a determinant of successful
190 Douglas-fir: The Genus Pseudotsuga
Sorensen and Campbell (1985) reported a seed was shown not to be correlated with seed size.
sophis- ticated study in which Douglas-fir seed
weight was artificially altered on each of several
seed trees by differentially bagging cones during
seed and cone development. The procedure
resulted in increasing average seed weight from
10.70 mg to 11.85 mg, an increase of 10.7%. The
treatment did not, however, increase the resultant
seedling height from 29.9 cm to 31.1 cm. They
reviewed literature reporting incon- sistent
relations between seed weight and resultant
seedling size, and offered the following possible
reasons (p. 1111):
• maternal genetic factors that affect seed
size differently than they do growth
• interactions between seed weight and genetic
differences in seedling growth habit
• influence of test environment on effects of seed
weight on plant size
• competitive effects among seedlings
Sorensen and Campbell (1985) also noted that a
3- to 4-year difference in 2-year-old seedling
height was associated with 10% differences in
seed weight, and suggested that "increasing seed
size compares favor- ably with other nursery
treatments for enhancing growth" (p. 1113). They
did not appear to have made further progress in
this area, however.
In a later paper, Sorensen and Campbell (1993)
suggested that the relation between seed size and
subsequent seedling vigor has two components:
environmental seed weight, an effect that
diminishes over time, and genetics, which reflects
the fact that the same genes that cause greater
seed growth are also reflected by greater
vegetative growth in the parent and in the
seedling, and that these effects do not necessarily
diminish with time.
According to St. Clair and Adams (1991),
In conifers, variation in the average weight of seed
col- lected from different female parents is a
consequence of three factors: the mean diploid
genotype of the embryo, the mean haploid genotype
of the megagametophyte, and the environmental
effects common to the mother tree. The latter two
factors are maternal effects. Evidence suggests that
family differences in seed weight and other seed traits
largely result from maternal effects. (St. Clair and
Adams 1991, p. 987)

Based upon 35 families, the percentage of filled


St. Clair and Adams (1991) also noted that was taken to ensure that the seed response
“seed size was only weakly correlated reflected any possible effect of collection date and
positively with seedling weight” (p. 993) and not of processing procedures. Considerable
that this correlation may be ex- pected to variation was found among trees, but the highest
diminish with time. They reviewed several germination and
papers indicating that, for Douglas-fir and
coniferous species generally, any correlation
between seed size and seedling size was either
weak or transitory, and they suggest that this
finding may be influenced by environmental
factors, stratification, and maturity on seed
characters (p. 993).

Seed development
From the moment that fertilization results in a
zy- gote, a complex series of biochemical
events occurs in the various tissues of the
coniferous megagame- tophyte, the end result
of which is the mature seed. In a lengthy and
detailed description, Bowley and Black (1994,
pp. 35–140) discussed this sequence for
angiospermous plants. As Edwards (1980)
noted, "according to Nitsch (1965) there are
four phases of growth and development in
seeds, which may be described briefly as (i)
cell initiation and multiplica- tion within the
flower bud, (ii) pollination, pollen growth and
fertilization. (iii) cell enlargement in the fruit
and cell multiplication in the seed and (iv)
maturation and finally senescence” (p. 627).
That work was based on research with
gymnospermous plants; much less
information is available for gym- nospermous
species: “For conifers, opening of the cones
on the tree and seed shedding usually signi-
fies that the seeds are mature.” (Edwards, p.
627).
Perhaps the first study of Douglas-fir seed matu-
rity was that of Finnis (1950). Cones were
collected weekly between July 19 and
September 28. The data showed that cone
specific gravity and seed weight were not
related to seed maturity, and that germina-
tion for both stratified and non-stratified seed
rose sharply for collections of the week of
August 16–23. The increase in embryo length,
but not changes in cone color, appeared to be
correlated with seed ma- turity, as measured
by germination. Allen (1958b) worked with
seed collected at 2-week intervals be- tween
August 15 and October 10. Every precaution
Chapter 7. Seeds 191
Although establishing the
most rapid germination occurred with seed from
the last collection. Ching and Ching (1962)
collected cones from April 6 until September 9.
The data showed these trends: respiration,
moisture content, and specific gravity of cones
decreased with increas- ing maturity; weight of
cone and seed, length of em- bryo, seed
production per cone, and seedling vigor (from
germinated seeds) increased with increasing
maturity: however, none of the foregoing defined
the point of maximum maturity. Ching and Ching
(1962) also noted the following for cones:
From the chronological changes, five distinct
develop- mental stages were indicated; receptive—in
the middle portion of April; enlarging—from late
April to June; filling and seed developing—from June
1 to August; maturing—from early to late August; and
drying—from late August to early September.
Browning of the cone bracts and length of embryo
more than 90% of the embryo cavity in seed will be
good practical indices for maturity of Douglas-fir
cones. (Ching 1962, p. 29)

(Note that cone bracts should be brown, the


embryo 90% of embryo cavity.)
Rediske (1961) studied changes in the
biochem- istry of Douglas-fir seed as it matured.
He found that reduced sugar levels of immature
and mature seed (22 mg/gm and 13 mg/gm,
respectively) were a good measure of seed
maturity. He also found small differences in
starch, soluble protein, and nitrogen with maturity
state, as well as differences in crude fat associated
with decreasing seed moisture of the cone. He
agreed with Ching and Ching that seedlings from
immature seed were less vigorous than those from
mature seeds.
Generally, maximum seed maturity and
germina- bility occur when seed is shed, but
maturity varies between trees, stands, and
between cones on the same tree (Kolotelo 1997).
The megagametophyte tissue of mature seed does
not shrink when seed are bisected (Eremko et al.
1989, p. 26). This finding is in agreement with
that of Harrington (1972, p. 148), who noted that
mature seeds can generally withstand damage
from drying. In a review of previous studies,
Tanaka (1984, p. 28) noted that color and firmness
of embryo and megagametophyte, seed wing and
cone scale color, and loss of cone and seed
moisture may all be used to estimate seed
maturity. The studies discussed thus far have been
concerned with physi- ological maturity of seed.
month before seed fall; however, Tanaka (1984, p.
exact time of maturity is difficult, it is 29) noted that the method is not widely used in
important because—as Harrington (1972, p. the Pacific Northwest, probably because of a
148) pointed out— immature seeds are not higher risk of poor germination and reduced seed
resistant to stresses such as drying, and do not yield. Silen (1958,
store well. Also, according to Harrington
(1972), “physiological maturity, even if not a
precise point in the life of a seed or not
precisely determinable, is still of extreme
importance since it marks the moment when
the seed begins to age. At this point, the seed
has its highest vigor, there after declining to
senescence and eventually no longer able to
germinate” (p. 152).
Several studies of the effect of cone
harvest at various dates, in effect at different
stages of seed maturity, on subsequent seed
germination show clearly the importance of
seed maturity on sub- sequent seed vigor.
Rediske (1969) studied seeds collected on
four dates. Seeds from the earliest date (July
23) were clearly immature, and only 1%
germi- nated. Seeds collected later were more
mature, had a higher germination rate, were
heavier, and produced larger seedlings.
Sorensen (1980) worked with seed collected 6
and 2 weeks before natural seed fall. The
seeds collected later had higher germination,
were 19% heavier, and produced larger
seedlings than the early-collected seeds. All
stratification benefited late- collected seeds, but
immature seeds were adversely affected by
stratification periods greater than 30 days.
Olson and Silen (1975) collected cones
between August 12 and September 14, with
the following results: “Cones, seeds, and
seedlings from 70 of 309 parent trees were
collected too early, resulting in poor cone
yields, reduced seed weight, poor ger-
mination, bacterial susceptibility, reduced
seedbed density, and greater expenditure of
time”(p. 11); and “seed weight and nursery
germination continued to improve steadily for
collections made even during
the last two weeks before seed fall” (p. 10).
Theisen (1980, p. 8) reviewed a range of
reports for conifers (including Douglas-fir),
which indicate that properly stored cones
collected before seed fall can yield fully
mature seeds. Silen (1958) reported that
Douglas-fir cones stored in damp peat moss
can yield mature seed from cones collected a
192 Douglas-fir: The Genus Pseudotsuga
p. 413) observed that fully mature seed weighed 14 days of germination, and that the reaction products
more than 8 mg, and suggested that seed weight of lipases increased. In an earlier paper TM
be a measure of seed maturity. Harrington (1972)
noted that “the most generally accepted measure
of maturity is the time when the seed has reduced
its maximum dry weight” (p. 151). Edwards
(1980) found that "hormonal levels appear mostly
to be associated with meristematic activity and the
high- est levels are found in immature seeds" (p.
627). No references were found relating plant
growth regula- tors to maturity in conifer seeds,
however.
The presence of starch in cone or seeds did not
provide a definitive measure of seed maturity
(Ching and Ching 1962). In an early study of
Douglas-fir seed metabolism before and during
germination, TM Ching (1959, p. 554) noted that
the first sign of germination of a seed is the rapid
increase of respiration that often starts within 1
hour after the commencement of water imbibition,
and that “seeds
soaked in H2O2 had a higher respiratory rate than
those soaked in water” (p. 560). Other studies
(Ching 1963a,b) noted that mature Douglas-fir
seed had
a high fat content (in common with 90% of seeds
studied by Bradbeer 1988, p. 24), which was used
during germination (Ching and Fung 1963, p.
551). To investigate this further they initiated a
study of labeled glucose. They found that the dry
weight of seeds and cone scales increased, while
their mois- ture content percentage fell with
increasing seed maturity. With increasing
maturity, the uptake of glucose and the respiratory
rate declined markedly, but the seed respiratory
quotient (RQ) remained above 1 for the entire
study.
The rate of fat synthesis from the labeled
glucose increased with increasing seed maturity,
so that increasingly mature seeds contained (p.
553) 40% fatty substances, 30% nitrogenous
components, 20% fiber, 4% other carbohydrates,
and 4% minerals. Seeds increased in dry weight
from 3.9 mg to over 12 mg during the study, and
decreased in MC from 84.6% to 59.4%. These
changes did not define the point of maximum
maturity, however. Working with mature seeds,
TM Ching (1968, p. 482) noted that lipids made
up 35% of seeds. Her data showed a significant
decrease in protein body-nitrogen during the first
Ching (1963a) noted that lipids are the true supported those of Ching (1963). He showed that
reserve of Douglas-fir seed, used during seeds were completely ripened after 10 days of
germination. She also noted that “the stratification, lipid reserves largely disappeared
glycerides are probably the major component during strati- fication, and seeds’ respiratory
which provides energy through oxidative activity increased after ripening from the low
degradation and carbon fragments for level of dormant seeds. Unfortunately, neither
synthesis of cellular material in future Ross nor Ching presented a quantitative measure
growth” (p. 231). She also found that “there of “after ripening.” Ross found that stratification
was no apparent change in total nitrogen facilitated a faster mobilization of storage
content of various stages of germination” (p. reserves and suggested that there “was a block
231). In a second report, TM Ching (1963b) preventing lipid breakdown in dormant
found that “total fats decreased rapidly with seeds” (p. 272).
germination from 36 to 12%” (p. 724), and Sorensen (1999) showed that drying storage of
that “a preferential utilization of linoleic acid Douglas-fir seed at 3°C or −12°C for up to 32
in glycerides and a preferential increase of weeks slightly reduced seed dormancy, whereas
linoleic and palmitic acids in phospholipids longer storage of up to 2 years at 12°C did not
were clearly demonstrated” (p. 728). In a later affect germi- nation rate or total germination.
paper, TM Ching (1966) noted that “lipids, Gosling et al. (2003,
proteins and reserve phosphorous com- p. 244) compared dormancy release during drying
pounds in the gametophyte were utilized for after ripening and during prechill, and noted that
the synthesis of carbohydrates, structural after-ripening in dry storage generally took place
components and soluble compounds in the most rapidly at low moisture content (in the range
seedling . . . data presented in this paper of 5% to 20% MC fresh weight). The moisture
indicate that the metabolic events of content fresh weight of the seeds Sorensen used
germination in gymnosperms are similar to was 8.46%, so a greater effect of storage on seed
those characteristic for angiosperm seeds” (p. after ripening than he reported.
313). The findings of Ross (1969) largely
Chapter 7. Seeds 193

Seed processing (damage) Although mechanically damaged seeds may appear


Although Douglas-fir seed frequently falls from Allen (1958) supplemented his samples from seed
the top of trees that are 60+ m in height, the wing, processing with trials in which measured blows were
which acts as a miniature helicopter, reduces the administered to individual seeds. The results from
impact of landing. About five decades ago, these trials demonstrated that relatively minor impacts
frequent erratic data of germination tests inspired could crack seed coats, and that where a single impact
some scientists to research the ability of the resulted in no measurable damage, repeated mild
supposed “tough” seed coat to withstand the blows could destroy seed viability. Further, this work
stresses common to the cone and seed processing showed that any crack in the seed coat resulted in
scenario used by commercial seed processing reduced viability, discolored radicle, and generally
facilities. In two publications (Allen 1957b, poor germination, and for even slightly damaged seed
1958a), seeds were sampled at stages of their stored poorly, dewing- ing was identified as a major
journey through the several operations necessary cause of seed damage. More modern equipment,
to remove seeds from the cone and to separate however, largely reduced the destructive impact of
seeds from extraneous material released from the dewinging (Edwards 1985,
cones with the seed. These trials demonstrated
that, though the actual reduction in seed viability
resulting from processing, particularly dewinging,
was frequently a function of seed moisture content
and maturity, the overall pattern was that seed
processing equipment generally reduced seed
viability.
normal, there may be subtle, detrimental is extremely important because it marks the
effects on seedling vigor. The physiological moment when the seed begins to age. The
basis for this loss of vigor is poorly external changes in Douglas-fir cones during
understood; causes may be physiological seed maturation are summarized in Table 7.7
deterioration triggered by impaction or (Ching and Ching 1962). The seed was mature
physical damage resulting from cryptic, by about August 12, although changes in cone
microscopic breaks at crucial spots in the appearance were too gradual to afford a precise
seed. Other researchers have examined the guide to seed maturity. One guide to
role of thigmomorphogenesis in reduced plant determining seed maturity is squir- rel activity.
growth (Telewski 1990, Telewski and Jaffe In late July or early August, squirrels frequently
1986). In conclusion, Douglas-fir seeds are frag- cut a few cones and then tear them to pieces,
ile and must be handled carefully to avoid leaving a pile of cone scales. This sampling
damage. continues until the seeds are mature, when the
squir- rels harvest many cones without cutting
Seed storage
them apart. Thus, when squirrels are actively
Harrington (1972) defined the initiation of harvesting cones,
seed storage as “the moment when the seed the seeds are mature.
is physi- cally mature” (p. 152). He also
noted that physical maturity may not be Table 7.7 External changes in Douglas-fir cones during the period
precisely determinable, but that the timing of seed maturation as indicators of maturity
p. 90). Leadem et al. (1990, p. 202) advised caution
Collection date Cone appearance
in order to avoid damage from excessive abrasion that may
result in poor germination. Barnes (1985) warned that worn July 23 Seed wing browning

rubber paddles on dewinging machines can cause seed July 30 Seed wing all brown
damage. Edgren (1968) noted that helicopter seeding August 5 Bracts yellowing
devices may damage seed. Stoleson and Hallman (1972) August 12 Bracts yellowish
reported data that supported Edgren’s conclusions. August 19 Bracts yellow to brown
Copeland and McDonald (2001, pp. 112–113) August 26 Cone greenish-yellow, yellow, or brown
noted that mechanical damage can affect agricul- September 2 Cone yellowish to brown, 0%–100% open on
tural seeds, resulting in cracked and broken seeds, different trees
baldheads, and other germination abnormalities. September
From 9 Ching 1962.
Ching and Cone brown, 0%–100% open on different
trees
194 Douglas-fir: The Genus Pseudotsuga
A second and perhaps more frequent volume of research dealing with the storage of tree
practical guide to seed maturity is the relation seeds was
between em- bryo length and the length of the
embryo cavity. When the length is at least 90%
of the cavity, the seed is mature. Immature seed
are generally fur- ther characterized by a milky
megagametophyte, whereas the
megagametophyte of mature seed is firm. A
discussion of seed storage may logically be
divided between factors affecting artificial
storage and storage under field conditions.
Storage under controlled conditions
Storage begins when the seed is mature, but seed
is often collected before cone and seed maturity.
Under these conditions, storage begins with the
seed in the cones—a procedure that has led to
conflicting data on the viability of seed after a
ripening period in the cone. Shea (1960) suggested
that fungi caused the loss of germination of seed
stored in cones. Rediske and Shea (1965)
demonstrated that, if cones were stored with more
than 40% MC or above 20°C, a great loss of seed
occurred after 16 weeks. Bloomberg (1969) noted
that “in general germinabilty losses after up to
120 days cone storage appear to be minor,” (p.
181) although he did find some diseased seeds in
the germination dishes. Lavender (1958), working
with seeds collected from 40 widely spaced trees
in the Willamette Valley, found that “storing
cones (up to 4 months at temperatures near 50°F)
does not affect the germinative capacity of
Douglas-fir seeds" (p. 8). Unfortunately, the
moisture content of these cones was not
determined. Ching and Ching (1962) suggest,
however, that for cones at the stage of maturity
described in the Lavender (1958) study, the
moisture content was about 15%.
A second form of storage under controlled
con- ditions is initiated after the cones have
been dried and the seed extracted, dewinged,
and cleaned. As Harrington (1972, p. 145)
pointed out, learning how to best store seed from
the harvest was essential in the shift from
hunting and gathering to the culti- vation of
crops. Early farmers had to learn how to guard
against high temperature and high humidity in
the stored seeds; the same problems faced
forest- ers. That the problems stimulated a large
particularly important because many forest Seed mois- ture was about 6%. Rediske (1967),
trees, including Douglas-fir, are characterized reviewing the literature, noted that high quality
by widely varying seed crops from one year to Douglas-fir seed
the next, so the only way to assure sufficient
seed for reforestation in a given year is to
successfully store them. Holmes and
Buszewicz (1958, p. 25) reviewed the
literature on storage environments for
temperate tree species and noted that, for
Douglas-fir, low moisture content (around
6%–9%) and low temperature (−18°C) pro-
vided the best storage conditions. Belcher
(1982) re- ported that “stratified Douglas-fir
seed can be dried to between 21% and 26%
MC and held at 3°C and stored for 10 months
without a significant decrease in
germinability” (p. 24). In their review of
forest tree seed, Edwards et al. (unpublished
note) stated that “optimum conditions vary
with species, but moisture levels between 6
and 9% (of fresh weight) and temperatures
around −18°C are widely used to maximize
seed longevity.” The general relation between
storage temperature and moisture content is
that, at any given moisture content, seed
viability de- teriorates faster as temperature
rises (within limits), and the lower the storage
temperature, the greater the tolerance to high
moisture content. Thus, stor- age temperature
is more important when moisture content is
high. Refrigeration equipment is expensive to
install and maintain, so paying close attention
to moisture content, which can be controlled
more economically, makes sense.
In the Pacific Northwest, seeds are commonly
stored at −18°C, but the British have found
0°C to be equally effective (personal
communication, AG Gordon, 1978). In two
reports, Barton (1954 a,b) found that, after 3
years, seed stored at −18°C retained higher
viability than seed stored at −4°C or −11°C,
but that the seed also had the lowest MC,
10%. Seeds stored at higher temperatures had
higher MC (15% and 16%), so the MC likely
interacted with storage and temperature.
Subsequent trials showed that seed at 5.8%
MC retained greater viability at both −18°C
and 5°C than did seed with 13.6% MC at the
same temperature. Allen (1957) reported that
Douglas-fir seed stored better at −17.8°C than
at 0°C or at room temperature for 7 years.
Chapter 7. Seeds 195
was about 7%. Further, no difference was shown
could be stored for at least 8 years at 6% to 10% between the viability of seed
moisture content and −17.8°C temperature
without loss. Schubert (1954) reported that
Douglas-fir seed stored at 5°C (no mention of
moisture) showed 66% viability after 6 years and
31% after 16 years. The germination tests were
conducted in greenhouses without
environmental control. Work reported by
MacMorran (1946) demonstrated that seed
stored at 2°C to 4°C for 3 years retained its
viability better than did seed stored at room
temperature, regard- less of whether seeds were
maintained in sealed or open bottles. A later
paper by Rudolph (1952) showed that seed
maintained its full viability for 4 years when
stored at 5°C in sealed containers. He
emphasized that, if seeds were to be stored for
pro- longed periods, they should have no more
than 7% MC when placed in storage. Sorensen
(1999, p. 96) working with Douglas-fir seed at
8.49% MC, found that seed stored at 3°C
deteriorated, and concluded that long-term
storage should be at subfreezing temperatures
(−10°C to −20°C).
One interesting variation on seed storage trials
was reported by Allen (1962a,b,c). He
presented data describing the effects of canning
Douglas-fir seed with and without vacuum (20
inches) and storing it at room temperature and
70% and 7.1% MC. Germination tests
demonstrated that vacuum storage could not
substitute for low-temperature storage. This
variation is, perhaps, occasioned by difference
in seed maturity and processing, both of which
can influence the effects of storage (Schubert and
Adams 1972). But for mature, undamaged seed,
low moisture content (6%–9%) and low
temperatures (below freezing) are apparently
most effective in maintaining seed viability.
In contrast to the long-term storage described
above, a short-term storage experiment by
Lavender (1954) demonstrated that exposing seed
to 30°C temperature and 93% relative humidity at
10°C for 11 weeks did not reduce seed
germination. A second study (Lavender 1958a)
showed no superiority for storage at −17°C over
that at 0°C or at the uncon- trolled temperatures in
an unheated warehouse during 6 months of the
fall, winter, and early spring, when seed moisture
seed did not retain its germinative capacity for
overwintered under field conditions and any of the more than a year under natural conditions in the
above storage regimes. western white pine region of Idaho. Finally, Isaac
A third study (Lavender 1958b) was (1943) reviewed
designed to determine whether seed viability
was reduced when cones were not extracted
shortly after picking. Populations of cones
collected from 40 widely sepa- rated trees in
the Willamette Valley were stored in an
unheated warehouse for 0, 2, and 4 months
before drying and extraction. No reduction
was found in seed vitality, as measured by
germination of seeds in greenhouse flats. All
the data discussed above are compatible with
the concept that moisture is essential for the
hydrolysis of seed substrate necessary for
germination and growth (Koller and Hadas
1982) and that respiration that uses seed
energy reserves increases with temperature.
Accordingly, long-term preservation of seed
substrates is best achieved under cold,
desiccating conditions. Schubert and Adams
(1971 pp. 50, 52) recommended below
freezing tem- peratures and 4%–8% MC.
Although substantial research has been
concerned with the storage of Douglas-fir
seed, data describing the effects of moisture,
temperature, seed maturity, and processing—
and their interactions over a period of at least
10 years—are lacking.
Storage under natural conditions
Obviously, nearly all of the above work dealt
with storage under controlled conditions. But
substantial effort has dealt with a second type
of storage: that is, storage under natural
conditions.
Early in the 20th century, Douglas-fir seed
stor- age in the duff was controversial. In
three papers, Hofmann (1917, 1920, 1924)
argued that the pattern of natural regeneration
after harvest or burning of Douglas-fir stands
could best be explained by hy- pothesizing
that it originated from seeds stored in the duff.
But he presented no data to demonstrate that
was possible. But Isaac (1935), using seed
stored for various periods under various
natural conditions, demonstrated that
Douglas-fir seed had a maximum life of 1
year in forest duff. Later work by Haig et al.
(1941) demonstrated that interior Douglas-fir
196 Douglas-fir: The Genus Pseudotsuga
the results of 9 years of trials with Douglas-fir upon seed surviving on the forest floor more than
seed stored in the duff under a variety of a year (Isaac (1943, p. 26).
conditions, finding conclusive "that Douglas-fir Garman (1955) noted early Canadian
seed does not commonly retain its viability for research demonstrating that Douglas-fir seed
more than a year after it ripens. There may be either germi- nated or died under natural
conditions in nature under which Douglas-fir seed conditions during the first year after it was
is germinable for more than a year, but the produced. Garman and Orr- Ewing (1949, p.
evidence is now very strong that the amount is too 15) found, however, that only 12% of the total
small to be a factor in reforestation, and forest germination of stratified Douglas-fir seed sown
managers should not count in May occurred during the first summer.
8. Seedlings
Denis P. Lavender

S
hoot growth in Douglas-fir and many other undergoes significant changes
conifers (e.g., spruce, fir, pine) is
indeterminate in the first year, but
determinate thereafter.
In many species, buds contain primordia for all
the leaves that will develop the following season.
Species that produce buds that contain, in
miniature form, all the growth that they will
accomplish in the subsequent year are called
determinate (restricted in growth). Species that do
not produce buds, or have buds that contain apical
meristems capable of ini- tiating further leaf
primordia and internodes (e.g., hemlocks), are
termed indeterminate.

Seedling Dormancy
Dormancy (from the French verb, dormir, to
sleep) is a term that refers to the vegetative buds
and tissues of a perennial plant. The remainder of
the plant does not have dormancy, but is strongly
influenced by this stage of bud physiology.
Dormany differs from cold hardiness, which
frequently develops at a similar time for Douglas-
firs and is systemic for all plant tissues. In
general, dormancy develops before the onset of
weather unfavorable for growth and represents a
survival mechanism, wherein growth is
suppressed for stress resistance.
The classic definition of dormancy (Doorenbos
1953, p.1) is “a tissue predisposed to elongate
does not do so.” A woody plant is generally said
to be “dormant,” by common usage, when buds
have formed on the terminals of shoots. With
many tem- perate plants, the dormant period may
extend from mid-summer until sometime in the
following spring, a period that can be more than
75% of the annual growth cycle. Although the
external morphology of the plant shows little
change during this time, the growth physiology
during this period, which govern the The first approach defines the annual growth
response of the plant to the environment. cycle of Douglas-fir as consisting of two
The dormancy phase of seedling contrasting states: (1) a period of active shoot
physiology is difficult to discuss because elongation, gener- ally from late March until mid-
researchers have inves- tigated it with August, wherein bud break and subsequent stem
dramatically different approaches, because the growth are accomplished by the elongation of pre-
subject has engendered a bewildering array of formed initials; and (2) dormancy, a period,
terminology (Lang et al. 1985), and because including bud set, of appar- ently no growth from
methodology involved has resulted in data mid-August until March. This pattern is typical of
that are frequently not comparable (Lavender determinate gymnosperms, as discussed below.
1991). The two major approaches to the study As noted in the introduction, there are a number
of dormancy in Douglas-fir are (1) a study of of reviews that discuss the subject of dormancy,
the morphology of the terminal shoot from so this section will refer primarily to these sources
bud break until and includ- ing bud set and the and to individual reports that present data
growth responses to defined environments; particularly germane to Douglas-fir.
and (2) a study (largely performed by Prof Sarvas (1974) proposed at least two stages of
J.N. Owens and associates) of the anatomy of dormancy: (I) the “chilling” period, (II) separated
apical meristems and adjacent tissues, by definite cytological events. He considered the
including cellular activity and its Dormancy I period to function as a mechanism to
biochemistry. set the physiology of a plant to the zero point of

197
198 Douglas-fir: The Genus Pseudotsuga
Dormancy II, i.e., bring all plants to an equal weeks of mild days with 9-hour photoperiods, fol-
state of readiness to utilize heat to initiate lowed by 4, 8, or 12 weeks of 9-hour days at 5°C,
spring growth. Sarvas suggested that there is a or were placed directly into 9-hour days at 5°C
distinct difference in Dormancy I and II, albeit from the natural conditions obtained in August.
the dividing line is dif- ficult to define. In After chilling, the seedlings were maintained with
contrast, Campbell (1978) argued that for 12- hour photoperiods at 20°C to evaluate bud
Douglas-fir, at least, dormancy is a period of break. The results showed clearly that the short-
transition “with potential developmental rates day (SD) treatment prior to chilling was essential
changing continuously in response to cool- for vigor- ous growth after chilling. Similar trials
season environmental stimuli” (p. 20). (Lavender and Wareing 1972) demonstrated that
Most of the research dealing with dormancy the sequence of long days-chilling resulted in 13%
has been discussed earlier in this paper, i.e., the mortality, as opposed to no mortality for SD-
effects of environmental factors in slowing shoot treated seedlings. Similarly, Jacobs et al. (2008)
elongation and initiating bud germination and the showed increased cold hardiness and root growth
environ- mental requirements for bud break. for Douglas-fir seedlings hardened under SD
These studies have introduced the concepts of regimes. MacDonald and Owens (2010)
summer dormancy or quiescence, rest or winter recommended a 3-week SD period for coastal
dormancy (Romberger 1963), and post-dormancy Douglas-fir (var. menziesii) seedlings after
or quiescence, all based upon the response compar- ing the effects of different SD treatments
(usually short-term) of a plant to favorable on bud development, bud endodormancy, and
environmental, i.e., warm tempera- tures and long morphology of first-year containerized coastal
photoperiods. This work has left a somewhat Douglas-fir (var. menziesii) seedlings in the
amorphous period between bud set and bud burst nursery, together with seedling survival and
for which there are few published data concerning growth after one growing season in a common
either the physiology of the plant or the garden. On the other hand, Taylor et al. (2011)
environment most favorable for the development found no differences in field performance
postulated by Campbell (1978), other than, of between Douglas-fir seedlings treated with short-
course, low temperatures. The remainder of this day dormancy induction and those for which
contribution will be concerned with these dormancy was induced with conventional
references. moisture and nutrient stress.
Several unpublished and published As earlier noted, “dormancy” is defined by a
(Lavender number of terms, none of which are truly defini-
and Wareing 1972) studies with Douglas-fir seed- tive. The following two tables (Table 8.1 and 8.2)
lings suggest strongly that photoperiod response, list some of the variety of terms used historically.
in this species at least, may be more subtle in Table
nature than bud break. Two-year-old seedlings 8.2 presents the nomenclature suggested by Lang
grown in pots under natural conditions during et al. (1985) in an attempt to present more
spring and summer until resting buds were well consistent, definitive terminology. The above
developed in late August were then exposed to terms are not
either 3 or 6

Table 8.1 Historical nomenclature of dormancy phenomena.

Reference Approximate definitions and equivalence of terms for physiological dormancy


Dormancy imposed by the environment— Dormancy imposed by agents or conditions Dormancy maintained by agents or
no internal control within the plant, but outside the dormant conditions within the organ itself
organ
Pro-rest middle rest after rest (Vorruhe,
Johannsen (1913) Forced idleness (Erzwungene Untätigkeit)
Mittelruhe. Nachruhe)
Doorenbos (1953) Imposed dormancy Summer dormancy Winter dormancy
Samish (1954) Quiescence Correlated inhibition Rest

Romberger (1963) Quiescence Correlated inhibition Rest


Chapter 8. Seedlings 199

Table 8.2 Ecodormany, paradormancy, and endodormancy.

Ecodormancy Paradormancy (later ectodormancy) Endodormancy


Regulated by physiological Regulated by physiological factors inside the
Regulated by environmental factors
factors
outside the affected structure affected structure

Examples Temperature extremes


Apical dominance Chilling responses
Nutrient deficiency
Photoperiodic responses Photoperiodic responses
After Lang et al.
Water stress
(1985).
Owens and Molder (1973) presented a micro-
scopic description of buds throughout dormancy.
without some criticism (Salisbury 1986 and They offered evidence that the vegetative apices pass
Junttila 1988) and certainly do not correct the
deficiencies pointed out by Lavender (1991), but
they have been in use by The American
Horticulture Society for 20 years, so we will
follow the above.
Webber et al. (1979) defined dormancy in
Douglas- fir as follows: “the period between
formation of a terminal bud (mid-July) to the
flushing of buds and beginning of spring growth.
In this period, Samish (1954) has described four
distinctly different physi- ological states:
quiescence, preliminary rest, midrest, and after rest
(each defined by growth response in a favorable
environment)” (p. 536). In Douglas-fir, these
states have been established from physiologi- cal
trials largely concerned with response to photo
period (Lavender et al. 1970).
Environment and dormancy initiation
By definition in Table 8.2, “ecodormancy”—
which in Douglas-fir, extends from mid-August
until late September—is imposed by the
environment. Vegis (1964) suggested that the
environment stimulating dormancy is that which
occurs prior to potentially damaging weather.
Accordingly, Douglas-fir gen- erally initiates
dormancy in response to the envi- ronment in
August, i.e., shortening photoperiods and dry
soils. Several reports have suggested that dry soils
are a major cause of dormancy (Blake et al. 1979;
MacDonald 1996; MacDonald and Owens
1993a,b; MacDonald and Owens 2006), while
others have found shortening photoperiods to be
gener- ally associated with dormancy initiation
(Lavender 1962, Lavender et al. 1968). Some
reports have indi- cated that low temperatures
may delay dormancy (Lavender and Overton
1972).
through five stages each year: dormancy
(November- March), early bud-scale initiation
(April-May), late bud-scale initiation (May-
June), early leaf initia- tion (July-August), and
late leaf initiation (August- October). These
stages are based on anatomical and
biochemical measurements which differ
strongly during the annual cycle and which
are more de- finitive than the morphological
descriptions more commonly used to define
dormancy and active shoot growth.

Dormancy and physiological response


In the detailed discussion of the apical
meristem described on the previous page
(Owens and Molder 1973) there is little
correlation between the stages of dormancy
and the dormancy phases described by Lang
et al. (1985). The concept of dormancy, as
described by lack of mitotic indices by Owens
and Molder, is much shorter than that of Lang
et al. In as much as the mitotic index (MI)
defines cell division, Owens and Molder’s
concept may be more distinct. Grob and
Owens (1994), who define MI as “a mea- sure
of the percentage of cells undergoing mitosis
at the time of fixation,” note the following,
however:
Interpretation of MI data requires an
understanding of the factors responsible for
changes in MI, and the realization that MI does
not indicate changes in all cell parameters under
all conditions. More rapid physiologi- cal tests
are required to predict seedling performance.
Cytological methods such as MI and the ability
to resume mitosis under promotive conditions
(Grob 1990) may be more accurate and rapid
than tests such as days to bud burst. This is
because they measure one process, mitosis,
which is more closely related to biochemical and
molecular processes [during dormancy] than the
more complex process resulting in bud burst.
(Grob and Owens 1994, p. 480)

Plant growth regulators and dormancy


For probably as long as dormancy has been
studied, scientists have had the
unsubstantiated belief that
200 Douglas-fir: The Genus Pseudotsuga
plant growth regulators (PGRs) have regulated membranes, i.e., the saturation and desaturation of the
this phenomenon. However, the following linoleate
summaries of PGRs and dormancy generally
agree that much of the evidence is erratic and
contradictory and has been generated by
insufficient methodology. Wareing and Saunders
(1971), Deyoe and Zaerr (1976), Doumas and
Zaerr (1987), and Webber et al. (1979) all noted
correlations between dormancy and levels of
PGRs, but Zaerr and Lavender (1980) and
Lavender and Silim (1987) agreed that varying
methodology made it impossible to establish
unequivocally the role of PGRs in dormancy.
While much of the foregoing research was
concerned with abscisic acid (ABA), the same
conclusions are true for indoleacetic acid and
other PGR compounds (Lavender and Silim 1987,
Saunders 1978). Perhaps the best conclusion is the
following, from Borchert (1991):
The hypothesis that bud dormancy in trees might be
caused by inhibitory plant hormones, such as abscisic
acid (ABA), was introduced 40 years ago, since the
level of growth inhibitors in extracts from dormant
Fraxinus buds declined during winter in parallel with
bud dor- mancy. Later, it was proposed that short
photoperiods cause an increase in the ABA content of
buds, while chilling reduces ABA levels and thus
enables bud break in spring. None of these hypotheses
withstood experi- mental testing [Lavender and Silim
1987, p. 171; Powell 1987b, p. 539], and the
following assessment of hormonal control of bud
dormancy, written 25 years ago, remains valid
[Romberger 1963]: “Our knowledge of endogenous
growth regulators (including morphogenetic receptor
pigments), and their interactions under various condi-
tions, is so inadequate that intelligent discussion of
the subject is not yet possible.” Indeed, neither shoot
growth periodicity nor any other aspect of plant
development involving correlations between organs
(e.g., apical domi- nance, flower induction, or tuber
formation) has been satisfactorily explained in terms
of hormone interactions [Davies 1987]. The genetic
and physiological control of morphogenesis is so
complex even in a relatively simple system such as
the isolated shoot meristem of tobacco [Meeks-
Wagner et al. 1989] that any attempt to deduce
hormonal control of shoot growth periodicity in
woody plants from crude correlations between
extracted hormones and shoot development appears
overly sim- plistic. There can be little doubt that plant
hormones are involved in the regulation of growth
periodicity. However, the complex interrelations
between environ- mental (photoperiod, drought, and
cold), nutritional, and hormonal factors remain to be
unraveled. (p. 240)

Erez (2000) offered an elegant discussion relat-


ing dormancy and cold hardiness in woody plants,
suggesting that the duration of dormancy is con-
trolled by changes in lipids in bud cell
and linolenate. Arora et al. (2003), who the “chilling” requirement of the bud, whereby
emphasized “the multiple and complex the bud must be exposed to temperatures of 5°C
nature of the dormancy phenomenon,” or lower for up to
explored bud dormancy in woody plants:
The path to endormancy induction is a
continuum, which in some plants begins as early
as budbreak in the spring. While it has been
tempting to explain bud dormancy on the basis of
hormonal regulation alone, dormancy is
controlled by numerous integrated plant
structures and functions (Crabbé 1994, Simpson
1990). Initial studies (e.g., Dennis and Edgerton
1961, Nitch 1957, Phillips and Wareing 1958,
Samish 1954, Wareing 1956) were followed in
the next 3 decades by a series of studies that
monitored endogenous levels of hor- mones
within whole buds, leaves, stems, cambium, and
root tissues under natural fall and dormancy-
inducing controlled-environment conditions. While
relatively easy to apply and measure responses,
many other problems are associated with
traditional exogenous application of hormones in
addition to degradation and differential responses
between the widely available commercial (±)
−ABA and the natural (+) −ABA (Wilen et al.
1996). (Arora et al. 2003, p. 912)

Certainly, the role of hormones in


Douglas-fir dormancy is not yet fully
understood. The PGR most commonly
associated with dormancy is abscisic acid
(ABA), previously referred to by Phillips and
Wareing (1958) as “dormin.” In unpublished
data, Lavender and Wareing (1969) noted that
ABA re- duced apical dominance of 2-year-
old Douglas-fir seedlings, but did not cause
dormancy. A report by Webber et al. (1979)
noted that the concentration of ABA was
highest in buds and needles in the autumn and
lowest in the same tissues just before bud
break. However, the authors do not assign a
causative role of ABA in initiating dormancy
in Douglas-fir.

Dormancy Breaking
The breaking of dormancy is at least a two-
stage phenomenon. The first occurs over a
period of time (probably 3-4 months for
Douglas-fir) wherein the factors responsible
for paradormancy are gradually dissipated,
generally by temperatures between 0°C and
10°C and the bud enters endodormancy. The
second phase, the elongation of the shoot and
con- current bud break, occurs as a result of
mild spring temperatures (5°C to 20°C).
The mechanism of the first phase is not
under- stood, but is generally referred to as
Chapter 8. Seedlings 201
relatively small mean temperature rise in the warmer
17 weeks before it will resume normal elongation portions of the Douglas-fir range during October to
(McCreary et al. 1990, Van den Driessche 1975, February
Wells 1979). It is difficult to assign a definite time
period under natural conditions because
interruption of the “chilling” by warm
temperatures (15°C to 25°C) may undo some of
the previous chilling, depending on the timing and
duration of the warm periods. This “chilling”
requirement, which is almost universal for woody
determinate perennials native to areas with frost
events in fall-winter, is not fully understood. More
recent reports, conducted with Douglas-
fir seedlings in exposed environments (Guak et al.
1998, Bailey and Harrington 2006), demonstrated
that Douglas-fir needs chilling; however, the
studies differ in that they employ air temperatures
rather than bud temperature and, as Chandler
(1957) stated, this can make a substantial
difference.
In a biochemically oriented review, Arora et al.
(2003) concluded,
Finally, while most work to date has focused on
hormon- al control of dormancy release, which, when,
how, and to what degree hormones are involved is still
uncertain, and evidence both supporting and refuting
various growth regulators can be found in recent
literature. What is clear is that, aside from more useful
hormonal localization studies and use of mutants and
transgenics, continued gross-analysis of hormone
presence or absence during dormancy release will not
enable definitive mechanisms to be tested. (Arora et
al. 2003, p. 913).

Climate Change and Chilling


As part of its cold hardiness, Douglas-fir has
evolved the requirement for a period of cool
temperature to break paradormancy. This
requirement has undoubt- edly saved the species
considerable frost damage. McCreary et al. (1990)
demonstrated that Douglas- fir seedlings grown
from either seed collected in an area with a
relatively warm winter or from seed collected
from trees growing in an area with a cold winter
break their buds more rapidly and produce more
vigorous shoot growth when chilled at a tem-
perature of 5°C than when chilled at temperatures
of 7°C or 9°C.
The significance of these data is that, while the
present climate of coastal North America has win-
ters sufficiently cold to satisfy the chilling
require- ments of endogenous Douglas-fir, a
(1983) predicted that winters with a mean
could well be sufficient to prevent this species temperature as much as 5°C above the present
from receiving its necessary chilling. aver- age would be within the expected range of
Therefore, the trees will die, either as a direct climatic variation after the year 2000.
effect of lack of chilling and failures of bud Accordingly, we may expect that poorly
break or by damage from bark beetles conditioned nursery stock will be
(Lavender 1989). For example, Copes (1983)
reported that a Douglas-fir seed orchard
established near the Monterrey Coast in
California demonstrat- ed very weak shoot
growth as a result of average winter
temperatures between 9.3°C to 12.2°C from
November to March.
Long-term weather records from stations
located in the Oregon Coast and Cascade
Mountains contain data from stations whose
mean temperatures in December, January, and
February are between 5°C and 8°C (Simonson
1963). Many papers cite data that predict a
mean global warming of 3-4°C within this
century. Furthermore, the majority of this
warming is predicted to occur during winters.
If, then, the mean winter temperatures of
forested areas below 300 m elevation in the
Oregon Coast Range and in parts of the
Oregon Cascades are raised by even those few
degrees, the average winter climate in those
areas would be too warm to satisfy the
chilling requirements of Douglas-fir and a
situation similar to that reported by Copes
(1983) would result. The findings of
McCreary et al. (1990) suggested that the
chilling requirement of Douglas-fir is not
influenced by the winter climate of the seed
source. If this is generally true for the species,
it may prove difficult to reduce the chilling
requirement through forest tree breeding
techniques.
Perhaps of more immediate concern to foresters
is the effect of a trend toward increasing
winter temperatures upon the success of
reforestation. The majority of the present
nurseries that grow Douglas-fir seedlings in
Oregon, Washington, and even British
Columbia are in areas that currently receive
only slightly more chilling hours each year
than are required by Douglas-fir seedlings.
Further, the methodology of harvest,
shipping, and planting forest tree seedlings
definitely impacts their abil- ity to respond
to chilling temperatures. Seidel and Keyes
202 Douglas-fir: The Genus Pseudotsuga
increasingly at risk in the coming years (Lavender a number of reports, Lavender (1985) concluded
and Stafford 1985). that stress resistance in Douglas-fir was minimal
in October through early December, while Van
Dormancy and Growth Potential den Driessche and Cheung (1979) found that
Lavender and Hermann (1970) and Lavender et seedlings lifted in early fall or May were more
al. (1970) demonstrated different degrees of sensitive to con- ditions in cold storage than were
growth dur- ing para-, endo-, and ectodormancy. those lifted during the winter. Hermann (1967)
They reported that no lateral or terminal buds found that cold storage of seedlings in November
were stimulated during paradormancy, and that and March was associated with lower survival
maximum growth was stimulated during the than when seedlings were lifted and stored in
change from parador- mancy to endodormancy. January. Mckay (1992) and Mckay and Mason
Root growth declined from October to February. (1991) found greater electrolyte leakage in
Lateral cambia did not respond to stimulus until seedlings lifted in the fall than in the winter; Folk
December and made maximum growth in et al. (1999) reported similar results. Common
January and February. opera- tional practice in the Pacific Northwest is
Owens (1967) presented a detailed description to begin lifting and storage in mid-December. The
of the growth and maturation of tissues after the date of lift and the duration of storage depend on
buds break dormancy. Although these whether the seedlings will be cold- or freezer-
development details are beyond the scope of this stored, and where or when they will be
book, the major points Owens made are as outplanted.
follows: Camm et al. (1994) reviewed some of the
• The first indication of increased activity research on the methods of cold storage and the
within the buds after dormancy occurs during physiological effects on seedlings, summarizing
mid-March. Positive staining for succinate as follows:
dehydrogenase appears at this time, first in
the apical meristematic region of the bud. Stress resistance is generally lowest in conifer
seedlings in a natural field habitat at times when the
• The dormant buds usually begin to expand seedlings are actively growing. Conversely, plants are
and initiate axillary shoots during the last most resistant to a number of stresses (frost,
mechanical, darkness, etc.) during those periods when
week of March (in the Victoria area).
growth is minimal or zero. In seedling nurseries, the
• The shoot and its leaves elongate rapidly and cycle is interrupted by fall or win- ter lifting and
push the bud scales apart, which results in bud subsequent cold storage where the plants receive none
of the clues of the natural environment.
burst during the second week of April. Cold storage of conifer seedlings is widely
• The shoot area grows rapidly during bud scale practiced in the temperate and boreal regions of North
America, as well as in Scandinavia. Successful
initiation, completed by mid-June and into management of this technique involves an
July. understanding of the ways in which plants respond to
changes in temperature and photoperiod in yearly
• Leaves are initiated rapidly during July
cycles of growth and dormancy. Operationally in
and August, and then more slowly until nurseries, moisture stress or length- ening nights in
nearly mid-November (during this period midsummer stimulate the formation of resting buds
the developing bud is considered to be in (Lavender 1990). Early fall conditions (mild
endodormancy). temperature and long nights) promote rest, or true
dormancy, in the apical meristem and initiate cold
• Maturation of the foliage over the growing hardiness in the entire plant. Late fall condition (low
season (after dormancy), although a truly temperatures and very long nights) terminate rest and
mature leaf is not formed until the tree maximize cold hardiness. In principle, seedlings put
into cold storage at this stage and protected from
becomes dormant. naturally occurring environmental variations over the
course of winter should remain cold hardy and able to
Stress resistance and cold storage break bud upon receipt of the appropriate heat sum
Although for many years cold, dark storage was after planting the following spring. In practice, cold
storage is used primarily to facilitate nursery and
thought to be neutral, later studies found that it planting schedules, and growers sometimes work
is stressful (Camm et al. 1994). After against, rather than with, the biology of the tree.
summarizing (Camm et al. 1994, p. 311)
In reponse to these and other reports and
others, additional studies have been
conducted on cold
Chapter 8. Seedlings 203

storage of Douglas-fir with and without a daily Douglas-fir that was lifted and stored in the fall
photo period (see Table 8.3; Lavender et al. 1970, was relatively low. They also found that the
Hermann et al. 1972). correlation between survival and root electrolyte
leakage was strong, and suggested that weather in
Cold storage with light
Britain did not allow Douglas-fir to proceed
There are few references to this storage regime, as normally through a dormancy cycle. O’Reilly et
it is limited primarily to Douglas-fir. Camm et al. al. (1999) found that the mitotic index of
(1994) reviewed several publications and Douglas-fir seedling shoots was at a minimum in
unpublished work that demonstrated the increased November-February, while that of roots was
seedling vigor and survival potential for seedlings erratic for this period, but generally high. Cold
stored with a short photoperiod. They emphasized hardiness was affected by weather, but was
that the work was done with photoperiod intensity maximal in November-February; seedlings lifted
too low to permit photosynthesis, and that the and directly planted survived well during late
greatest positive effect occurred when seedlings winter and spring. They emphasized that the
were stored in September- November, although dormancy cycle in northern Britain and Ireland
seedlings reacted positively in later winter. They differed from that in northwestern North America.
suggested that the results may be due to the
photoperiod on training circadian rhythms Table 8.3 Effects of a daily photoperiod during cold storage
(Lavender 1988). Douglas-fir seedlings lifted in upon the growth responses of coniferous seedlings.
October were stored at 20°C with root tem- Previous trials
peratures either 5°C or 20°C and with light or in Douglas-fir seedlings stored from mid-October to mid-November
the dark for 1 month, with survival as shown in and then placed in a growth-promoting environment.
Table Seedling mortality 5%
8.4. When the study was repeated with seedlings Photoperiod 8h
lifted in January, survival was 100%. Dark 55 h
Douglas-fir seedlings stored from mid-January until mid-April were
Cold storage without light then outplanted and mean date of seedling bud break tallied.
It has been generally felt that cold storage was a Date of bud break*
safe, neutral way to maintain seedlings between
lifting
and planting. But this procedure does not take into Mean Dark 8-h daily 16-h daily
account the endogenous rhythm of seedlings; con- photoperiod photoperiod
tinuous darkness at a constant temperature 50% June 4 June 2 May 26

produces an environment that does not entrain May 31 May 27 May 19

these rhythms. As a result, seedling vigor is Douglas-fir seedlings maintained from mid-January until September
in a constant 4°C environment
consistently reduced in cold, dark storage, which
Date of bud break*
subsequently results in reduced seedling vigor
Dark 8-h daily 16-h daily
after planting. Ritchie (1987) noted that cold photoperiod photoperiod
storage slows release from dormancy, and that September August 2 July 15
significant quantities of food reserves are lost
*Only seedlings with 16-hour daily photoperiod had a normal budbreak.
through respiration during storage. McKay rec- All the above seedlings were stored at a constant 2°C. Light intensity was 500 Lux.
ommended storing Douglas-fir between mid-January
and mid-March, and found that survival was
Table 8.4 Storage condition and survival for seedlings lifted in
better after storage at +2°C than after –2°C. October.
McKay (1992) found that Douglas-fir lifted and
stored in early fall
had lower survival than did later-lifted stock, and
that electrolyte leakage from fine roots was a stored in mid-December survived and grew well.
good measure of seedling vitality. Mason and McKay and Mason (1991) found that post-storage
Sharpe (1992) reported that Douglas-fir lifted and survival of
Storage Condition Survival (%)
Light (1 mo) Warm roots (20°C) 100
Cold roots (5°C) 95
Dark (1 mo) Warm roots (20°C) 90
Cold roots (5°C) 45
204 Douglas-fir: The Genus Pseudotsuga
Van den Driessche (1977) showed that Douglas- may be used to estimate dormancy, but their data
fir of any provenance stored poorly at −5°C to reflected more correspondence between fluorescence
−9°C. He found that “cold storage at 2°C in a characteristics and frost resistance than dormancy.
sealed plastic- lined paper bag, satisfied the Binder et al. (1997) noted that “chlorophyll
chilling requirement for bud flushing in two fluores- cence is a non-destructive and rapid
coastal provenances of Douglas- fir to the same assessment of in vivo photosynthetic activity.” It
extent as open nursery conditions” (p. 130). A follows, then, that for chlorophyll fluorescence to
comparison of European and North American identify discrete stages in seedling dormancy,
results demonstrated that the dormancy cycle in such stages must differ abruptly and sharply in
Douglas-fir in the northwestern United States is photosynthesis. We know of no data to support
strongly influenced by summer drought and cold this hypothesis. As Binder et al. (1997) noted,
in the winter, and is definitely different from that however, “before chlorophyll fluorescence can be
in Britain. Carlson et al. (1980) presented a figure widely used for forestry applica- tions,
that relates the various concepts of dormancy. standardization of techniques and fluorometer
They noted that “the expansion of preformed stem parameters are required” (p. 64). They
and leaf primordia occurs during the bud scale summarized a wide range of trials, all involving
initiation phase. Free growth can occur during the physical dam- age to seedlings, in which
rapid leaf initiation followed by a return to bud chlorophyll fluorescence could successfully
scale initia- tion. Vegetative bud set occurs at the estimate seedling performance. Perks et al. (2001)
end of scale initiation” (p. 371). presented a detailed summa-
Hawkins and Binder (1990) summarized ry of chlorophyll fluorescence characteristics and
research on the concept of mitotic index to that relationships:
date: Measurements of root growth potential (RGP) can be
pre- empted by using assessments of shoot
The wide operational use is probably due to its photosynthetic processes, under ideal conditions.
apparent complexity and lack of applied operational The positive rela-
publications. However this should not detract from the tionship found suggests that photosynthetic
test. There are sufficient data to suggest that M.I. reactivation is rapid after removal from cold storage
could play an important role in optimization of stock to conditions “ideal” for growth, and this result may
quality during the bridging phase (lifting to planting be of particular relevance in post-planting
hole) in conjunction with testing of seedling stress assessments, that are used for prediction of survival.
resistance. For example in Douglas-fir, M.I. should The findings also suggest that RGP may not
remain at or near zero for seven days prior to lifting necessarily predict future performance and field
and storage. (Hawkins and Binder 1990, p. 104) survival, as poor root growth was evident for plants
which established successfully. This reinforces the
One of the weaknesses of studies of the gross notion that R.G.P. provides unreliable estimates of the
mor- phology of the apical shoot, however, is that quality of cold stored stock and should not be used
they do not identify the stage of dormancy at a as a stand-alone test. (Perks et al. 2001, p. 233)
The relationship between measures of RGP and
given time. Techniques such as the OSU vigor PSII photochemistry offers the potential for a
test (McCreary and Duryea 1965) are acceptable, significant reduction in the time required to predict
but require several weeks. Interest in more rapid the ability of the plant to produce new roots, under
favorable conditions but the utility of such
evaluation of dor- mancy status has stimulated measurements in
research in the areas that follow. predicting seedling survival appears limited. Thus,
modulated fluorescence measurements have the
Chlorophyll fluorescence poten- tial to provide an “instantaneous” measure
that, with further parameterisation to take into
Hawkins and Lister (1985) discussed chlorophyll account seasonal variability, could be used to
fluorescence and the measure of the state of the identify and predict the vitality of stock,
pho- tosynthetic complex in seedling foliage particularly that previously subjected to cold
storage. (Perks et al. 2001, pp. 233-234)
(chlorophyll a and b carotenoid contents),
concluding that such data may well be correlated Dormancy and the Concentration
with phases of dormancy. The advantage is that
fluorescence measurements may be made non-
of Inorganic and Organic
destructively and rapidly. Binder and Fielder Constituents
(1996) suggested that fluorescence curves We found little material with reference to dorman-
cy and the concentration of inorganic and
organic constituents in the literature. Tables
8.5 and 8.6 are
Chapter 8. Seedlings 205

from Ketchie and Lopushinsky (1981, pp. 4-5)


Increased root activity was strongly correlated with
and represent values of root pressure exudate. lowered reducing sugar concentrations in seedling
Aspartic acid, glutamic acid, and glutamine were roots of the faster growing source. Sucrose and
the main amino acids in the exudates from all raffinose in- creased markedly during early winter and
were ap- parently converted to starch in spring prior
species. This finding is similar to the results to growth. Concentrations of reducing sugars, crude
reported by Barnes (1963) and is to be expected, fat, and protein nitrogen changed little with seasons.
since these compounds are the main amino acids (Krueger and
Trappe 1967, p. 192)
involved in xylem transport of N and in the
Starch in tops remained low in autumn and early
transamination process. Exudates from Douglas- winter, but increased rapidly beginning in March. . . .
fir also contained large amounts of leucine and A peak was reached by mid April, followed closely by
alanine. Grannel et al. (1990) found a constant a rapid decline. Subsequently, concentrations
increased moderately during June and July. (Krueger
level of 10% to 15% dry weight of non- structural and Trappe 1967, p. 198)
carbohydrates, which rose sharply with growth Lopushinsky (1980) and Ketchie and
initiation in April. (Krueger and Trappe 1967) Lopushinky (1981) examined the root exudates of
presented a detailed description of food reserves
2-0 Douglas-fir seedlings. Lopushinsky (1980)
in Douglas-fir seedlings and associated growth:
noted the following: “The fact that exudation
occurred from the bare-root

Table 8.5 Concentration of constituents in root pressure exudates from individual Douglas-fir seedlings.

Seedling Concentration (%)


number Sugars Amino acids Organic acids N K Ca Mg pH
1 0.20 0.01 0.01 0.011 0.016 0.007 0.001 5.4
2 0.12 0.01 0.01 “ “ “ “ 5.4
3 0.23 0.01 0.02 “ “ “ “ 5.4
4 0.21 0.01 0.03 “ “ “ “ 5.4
5 0.13 0.02 0.01 0.014 0.018 0.004 0.001 5.4
6 0.25 0.03 0.01 “ “ “ “ 5.4
7 0.29 0.03 0.01 “ “ “ “ 5.4
8 0.34 0.04 0.01 “ “ “ “ 5.4
a
S.E. ± 0.08 ± 0.01 ± 0.01 ± 0.001 ± 0.001 ± 0.001 - -
Composite b
0.18 0.02 0.01 0.011 0.019 0.005 0.001 5.4
a
S.E. = standard error
b
Combined sap samples from eight other seedlings.
From Ketchie and Lopushinsky (1981).

Table 8.6 Sugars and amino acids in root pressure exudates of conifer seedlings.

Species Sugars Amino acids


Engelmann spruce a
Aspartic acid, leucine, glutamine, glutamic acid, glycine,
serine Grand fir Glucose, fructose, sucrose, unknown 1b, unknown 2c
Noble fir Glucose, fructose Aspartic acid, asparagines, glutamine, glutamic acid, glycine, serine,
Lodgepole pine Glucose, unknown arginine, leucine
1b Ponderosa pine Glucose, unknown
1b

Pacific silver fir Glucose Aspartic acid, glutamine, glutamic acid, leucine
Aspartic acid, arginine, asparagine, alanine, leucine, glycine, serine,
Douglas-fir Glucose, unknown 1b,d, unknown 2c glutamine, glutamic acid
a
Sugar concentration was too low for identification.
b
Unknown 1 had the same rf value on chromatograms as rhamnose.
c
Unknown 2 had the same rf value on chromatograms as ribose.
d
Unknown 1 was found in the exudates of only one of eight seedlings tested.
From Ketchie and Lopushinsky (1981).
206 Douglas-fir: The Genus Pseudotsuga
Douglas-fir seedlings with completely subarized not conclusive evidence that it never occurs in those
species. The results of the present experiments sug-
root systems lacking any root growth clearly gest that the best opportunity to observe exudation in
demon- strates that exudation was not dependent field-grown conifers probably is during early spring
on active root extension” (p. 278). He concluded, following snowmelt because of a combination of de-
sirable factors including ample moisture supply, low
The reasons for the abundant and persistent evaporative potential, and high plant sugar content.
exudation observed in the previous experiments are (Lopushinsky 1980, p. 279)
not entirely clear. Apparently the combination of
healthy seedlings, minimum moisture stress as a
Roberts et al. (1991, p. 439) noted that a 30-
result of enclosing the seedlings in plastic bags, and kDa protein began to accumulate in the (terminal)
cold storage created condi- tions conducive to bud tissue in early November and that by late
exudation. Exudation immediately after detopping
normally occurs only in well-watered, turgid plants
November, this protein had reached its maximum
so moisture equilibration probably was a factor in level; it re- mained at this level throughout the
the present experiments. However, prolonged cold winter. The apical bud began to swell in early
storage of the seedlings probably was the key factor
April, and by the middle of the month, needles
enhancing exudate production. It is well known that
low temperatures favor starch-to-sugar transforma- were protruding from the bud scales. The levels of
tions in plants (Meyer and Anderson 1952, p. 385- the 30-kDa protein had declined to indetectable
386; Siminovitch and others 1953). An increase in levels in seedlings by early April. Concentrations
soluble sugars could influence exudation both
through stimula- tion of respiration-dependent ion of reducing sugars, crude fat and protein nitrogen
transport into the root xylem, and a lowering of the changed little with seasons. Starch intake
osmotic potential of the xylem sap because of remained low in autumn and early winter but
increased sugar content.
The present results not only demonstrate the abil- increased rapidly beginning in March. Peaks were
ity of these particular species to exhibit root pressure reached by mid-April followed closely by a rapid
exudation but also emphasize that the failure of some decline.
conifers to show exudation under certain conditions is
9. Cone and Seed
Insects and
Diseases
Denis P. Lavender

A
lthough, strictly speaking, cone and seed in- 1. Obviously this does not apply to Megastigmus.
2. The name “Megastigmus spermatotrophus," which appears in some
sects are not part of the biology/physiology of references, has been retained in quotations only.
trees, their impact on the production of viable
Douglas-fir seed can be considerable, as Mattson
(1978) noted: “Large fruit and cone crops
preferen- tially mobilize and utilize an
abundance of nutri- ents and photosynthates. In
doing so they reduce cambial, shoot, root and
leaf growth (Matthews 1963, Tappeiner 1969,
Kozlowski and Keller 1966, Kozlowski 1971,
Puritch 1972). Cone insects can drastically reduce
the amount of nutrients and photosynthates that
are allocated to reproductive structures because
they kill the flowers, conelets, and cones early in
their development before such structures have
utilized large stores of energy and
nutrients” (p. 339).1

Cone and Seed Insects


The first published reports that discussed insects
associated with Douglas-fir cones and seed origi-
nated in Europe in the late nineteenth and early
twentieth centuries. Early work in Europe (Wachtl
1893, MacDougall 1906a,b) focused on the genus
Megastigmus, while investigations in North
America were concerned with both this genus
(Crosby 1909, 1913; Rohwer 1913; Miller 1916)
and the genus Barbara (Cooley 1908, Miller and
Patterson 1916). Dr. Fritz
A. Wachtl first described Megastigmus
spermotrophus2 from insects that emerged from
seed he received in Vienna in 1893. MacDougall
(1906a,b), Crosby (1909), and Rohwer (1913) all
presented data describing this same insect.
Early reports concerning cone and seed and seeds. However, only 18 species were
insects in the Pacific Northwest included described as phytophagous, and only a scant half
those of Miller (1914 and 1916), Willis and dozen of this group are sufficiently wide spread
Hofmann (1915), and Hofmann (1924). Miller and numerous enough to have engendered
(1914) noted significant dam- age to conifer appreciable research. Accordingly, this discussion
seed by a number of unnamed insects in will be limited to the following insects, which
southern Oregon. Publications by most authorities (Keen 1958, Koerber 1960,
silviculturists (Willis and Hofmann 1915, Schowalter et al. 1985, de Groot et al. 1994, Meso
Hofmann 1924, Isaac 1943) detailing insect 1979, Miller 1986a, Miller and Ruth 1989)
damage to Douglas-fir seeds were primarily consider to have the most important impacts on
taxonomic until about 1950, when interest in Douglas-fir cones and seeds. Megastigmus spermot-
cone and seed insects increased markedly. rophus (Hymenoptera: Torymidae), Barbara
This paralleled increases in both reforestation colfaxiana (Kearfott) (Lepidoptera: Olethreutidae),
and in the development of seed orchards, Contarinia oregonensis (Diptera: Cecidomyiidae),
which produced expensive seed in the western Dioryctria abietella (Denis & Schiffermüller 1775)
United States. The great majority of papers (Lepidoptera, Phyoitidae), Leptoglossus
discussing cone and seed insects of Douglas- occidentalis (Heidemann) (Hemiptera: Coreidae)
fir appeared during this 40-year period from and three insects, whose damage is generally
about 1950 to 1990. minor, but which maybe im- portant in some areas:
Keen (1958) listed more than 60 species of insects Lepesoma lecontei (Coleoptera:
that have been reared from Douglas-fir cones

207
208 Douglas-fir: The Genus Pseudotsuga
Curculionidae), Choristoneura occidentalis Freeman abietella”: larger cones and short duration of
(Lepidoptera: Tortricidae) and Contarinia washing- flushing were
tonensis Johnson (Diptera: Cecidomyiidae).
Keen (1958) reported that,
the cones of Douglas-fir are infested with a great
number of species of insects, the most destructive
being a seed chalcid and two species of moths
[Bedard 1938]. In general about 10 percent of the seed
crop is destroyed by these insects, although in
different localities, and in different years the damage
may vary from practically nothing to 50 percent or
more of the seed destroyed. The insect damage is
usually heaviest in years when the seed crop is light
[Hofmann 1920]; and in any one year the damage is
usually worse in the warmer places [Hofmann and
Willis 1915]. (Keen 1958, pp. 26-27)

The following species have been cited as of


economic importance (Keen 1958, p. 27): Barbara
colfaxiana vars., Dioryctria abietella, Eupithecia
albi- capitata Packard, Eupithecia spermaphaga
Dyar, and Megastigmus spermotrophus Wachtl.
Johnson and Winjum (1960) concurred in the
importance of the first two insects and the seed
chalcid, but suggested the following species as more
destructive in the Pacific Northwest than the two
Eupithecia: Henricus fuscodorsana, Contarinia
oregonensis, and Contarinia
n. sp. The following sections will summarize
some of the key research relevant to Douglas-fir
seed and cone damage for several of the above
insects.
Kozak (1963) conducted a lengthy, detailed
analy- sis in British Columbia on the distribution
of and interactions among three major species of
insects that damage Douglas-fir seed: Contarinia
oregonen- sis Foote, Megastigmus spermotrophus
Wachtl, and Dioryctria abietella D. & S. The
study was based on more than 7,500 cones total
(taken from 93 trees in 1961 and 97 trees in
1962). Kozak (1963, p. ii) found high variation in
damage among trees for each insect species: "In
C. oregonensis this variation was significantly
related to the height of the trees and dates when
cones became pendent”; i.e., taller trees had
greater attack. For M. spermotrophus, “the
percentage of filled seeds and average cone size
of the trees were important”: greater attack was
associated with a higher percentage of filled seed;
smaller cones (with perhaps thinner cones scales)
had greater attack. Finally, “average cone size of
the trees and duration of vegetative bud flushing
were significantly associated with For D.
associated with higher damage (Kozak 1963, larva does not locate this point, it will perish.
p. ii). Data presented demonstrated that C. Mortality at this stage is generally high; as Hedlin
oregonensis was the most destructive and D. (1960) noted, only 55% of the larvae were
abietella was the least destructive to cones
and seeds of the insects studied. Insect-caused
seed loss varied widely with location and
year, from less than 5% to more than 90% of
Douglas-fir seeds. The findings showed that
D. abietella damaged 18.2% and 5.7% of cones
in 1961 and 1962, respectively (Kozak 1963).
Meso (1979) found that the greatest
percentage of losses occurred in years of light
to medium seed crops following years of
heavy crops; Kozak (1964) noted that the
greater the number of cones, the less insect
damage per cone. In 1983, C. oregonensis and
M. spermotrophus together destroyed 70% of
Douglas- fir seed in 17 seed orchards in
California, Oregon, and Washington
(Schowalter et al. 1985). Hall (1955) noted
that 97.8% of the seed was destroyed on the
Klamath National Forest; 91.6% on the Six
Rivers Forest, 53.2% on the Lassen; overall,
during the light cone year of 1954, 82.2% of
Douglas-fir seed was destroyed by insects.
Barbara colfaxiana
Barbara colfaxiana (Kearfott) insect was
apparently discovered in 1900 by R.A. Cooley.
Cooley prepared a bulletin, “The Douglas
spruce cone moth,” in which he identified it as
Cydia pseudotsugana Kearf. Keen (1953)
corrected the nomenclature to the present
Barbara colfaxiana.
The adult insects emerge from
overwintering pu- pae about the time the
pollen matures on Douglas-fir trees. The
female moth oviposits primarily in the late
afternoon and evening. No preference is
shown for levels in the crown, but exposed
cones are attacked more frequently than are
those shaded by foliage. The female deposits
her egg on the outer surface of the bract (in
areas of heavy infestation, several eggs may
be deposited on a single bract). The pearl-
colored egg is glued to the bract, where it
remains for a 2- to 3-week incubation period.
When the larva hatches, it must find the angle
between the bract and the cone scale where it
can force itself through the heavily pubescent
surface of the scale to begin feeding. If the
Chapter 9. Cone and Seed Insects and Diseases 209
cone moth larva develops normally with the parasite
able to establish themselves in cone scales. And, feeding internally until the cone moth larvae spins a
even if the larvae are successful in penetrating cocoon, at which point the parasite kills the
the cone scales, many more are killed by high
levels of pitch in the feeding tunnel. The initial
tunnel is very small, but becomes gradually
larger as the larva burrows towards the cone
axis, feeding primarily on cone scale material.
When the first seeds are destroyed, the larva is
small enough to burrow into the seed; later it
will consume the entire seed.
The insect passes through four instars by mid-
Ju- ly, and more than one larva may inhabit the
feeding tunnels in the cone without cannibalistic
behavior. The mature larva constructs a tough
cocoon, which becomes covered with pitch
adjacent to the cone axis, and pupates. The insect
remains dormant in the cone until the following
April or May, when most of the adults emerge.
However, Hedlin et al. (1982) found that a
significant proportion of the insects may enter a
diapause of 1 year or longer, and hence avoid
extreme competition for limited resources in years
of light cone crops. The findings of Hedlin et al.
(1982) are interesting in that they showed that
factors such as daily maximum and mean
tempera- tures during the period of larval
development were strongly correlated with degree
of diapause the fol- lowing year and with the level
of cone crops. When the year following larval
development had no cones on Douglas-fir trees, as
many as 80% of the Barbara pupae were in
diapause.
Although by the time the insect has reached the
fourth instar, one to three larvae may have
consumed much of the interior of the cone, there
may well be little external evidence of such
activity in cones of coastal Douglas-fir. If,
however, as many as five or six larvae are present,
the cones wither prematurely. In contrast, the
cones of var. glauca, which are somewhat smaller
than those of var. menziesii, commonly have
external signs of the presence of even a single
larva. Keen (1958) lists 12 hymenopterous
species as parasites of B. colfaxiana and
suggested that such parasitism, which can be as
high as 83% of the cone moth larvae, is a major
control of this insect. Hedlin (1960) noted that
Glypta evetriae Cush. females lay eggs in early
instar larvae of B. colfaxiana. The para- sitized
western North America, where it feeds on a range
host. Hedlin noted that only 2% of the cone of coniferous seed. Koerber (1963) presented a
moth eggs tallied on one tree produced detailed life history of this insect. The larvae have
larvae which lived to pupate, and that 92% five instars and feed on seeds during the summer
of the pupae were parasit- ized by Glypta.
As in the case of Megastigmus
spermotrophus, re- ports of damage caused by
Barbara colfaxiana vary greatly, although the
consensus of workers in the field is that this is
a major predator of Douglas-fir seeds. Kozak
(1963) was unable to find evidence of the
cone moth in coastal British Columbia, but
Radcliffe (1952) found that this insect
destroyed about 60% of the seed on
Vancouver Island, and that one larva per cone
would consume about 45% of the seed, three
would consume 75%, and more than three
insects per cone, about 100%. Hedlin (1974)
noted that B. colfaxiana was more damaging
in the interior of British Columbia than in the
wetter coastal regions. Schowalter et al.
(1985) found that the cone moth was
responsible for very little seed predation in
Oregon and Washington, but other workers
have identified significant seed loss due to
this insect.
Volney (1984) and Koerber (1960) presented data
demonstrating that although wild
populations of Barbara colfaxiana may be
limited by cone crop size, this insect can
compete more effectively for the food
resource than its competitors. Miller et al.
(1984) found that the degree of damage to
Douglas-fir seeds by B. colfaxiana in the
interior of British Columbia, but not on the
coast, was significantly related to the size of
the seed crop the previous year. However,
differences in cone crop size during the year
of dam- age and seed predation were not
correlated. Light crops following heavy
crops during the 33-year study were
generally heavily damaged. The authors
suggested that the fluctuating cone crop size
limits populations of B. colfaxiana in the
interior of British Columbia (see Roy 1960).

Leptoglossus occidentalis
An insect first described by Heidemann
(1910) is the western conifer seed bug,
Leptoglossus occidentalis (Heidemann). It is
widely distributed in the timber regions of
210 Douglas-fir: The Genus Pseudotsuga
months. The adults feed on the ripening seed in cones.
the fall, until cold weather when they find
shelter; they reappear in mid-May to feed on
year-old cones and seeds. The insect feeds by
forcing its stylets into the seed, dissolving the
contents with its saliva, and imbibing same. A
slight wound is the only evidence of the insect
attack, but the seeds are largely hollow because
the endosperm has been destroyed.
Zhang and Schowalter (1997) reviewed the lit-
erature, noting that, “feeding by the seed bugs . . .
causes an undetermined amount of seeds to abort
or not fill out” (p. 29) Given this, the “conifer
seed bug occurs sporadically and generally
destroys fewer seeds” than does Contarinia or
Megastigmus. However “at densities of 0.5 insects
per cone, seed bugs significantly increased seed
abortion—from 45 seeds in protected cones to 75
seeds in cones caged with seed bugs during seed
development in June.” The seed bugs “also
significantly increased the number of partially
filled seeds from 0.5 seed in control cones to 32
seeds in cones caged with bugs during July. Seed
yields were reduced 20 to 30% by seed bugs
feeding.” Zhang and Schowalter (1997) noted that
although the studies “indicate potential losses to
seed bugs, methods for measuring seed bug
populations in orchards are necessary to predict
impacts on seed yields” (p. 29).
Bates et al. (2000a) reported the following:
1. The major storage reserves of Douglas-fir
seeds are proteins and lipids, with the latter
making up the greatest proportion of the dry
weight of the seed. They further note that the
feeding procedure of L. occidentalis is
unknown and that it is difficult to determine
whether blank seed are the results of the L.
occidentalis feeding or some other cause, so
that absolute damage by seed bugs remains
unknown.

2. Seed lightly damaged by L. occidentalis had


>55% reduction in both lipid and crystalloid
protein resources and such seed showed a
>80% reduc- tion in germination.
3. Feeding by nymphs, adult males, and adult
females was similar and resulted in a reduc-
tion of full seeds in cones of »70% in a 2-
week feeding period, compared with coastal
Earlier feeding by nymphs resulted in a Torymidae (superfamily Chalcidoidea, order
threefold increase in the number of Hymenoptera) contains about 40 species, of which
unextractable seed. one-third are known to feed upon coniferous seeds
Blatt and Borden (1996) found that L. (Milliron 1949,
occidentalis oc- curred in patches and showed
a clonal preference; interestingly, seed losses
from this species were less than 5%. In a
subsequent study, Blatt and Borden (1998)
found that seed bugs did not feed on seed
infested with Megastigmus, although they
could feed on Megastigmus larvae. The
authors concluded that “the impacts of L.
occidentalis and M. spermatotro- phus are
segregated and additive” (p. 775). Lait et al.
(2001) reported the development of a
polyclonal antibody that can identify salivary
gland extracts of Leptoglossus occidentalis and
which is useful in iden- tifying light to severe
damage of Douglas-fir seed. Schowalter and
Sexton (1990) examined the pos- sible effects
of the timing of seed bug feeding on
Douglas-fir in seed:
Results of this study supported the hypothesis
that seed bug feeding on Douglas-fir seed has
different effects at different stages of seed
development. Seed bugs caused substantial seed
abortion, >50% greater than control levels . . .
during the early and mid-stages of seed de-
velopment. Significant increases in partially
filled seed resulted from seed bug feeding during
mid- and late stages of seed development.
While partially filled seed is detectable by X
ray and has been a recognized effect of seed bug
feeding on Douglas-fir seed (Koerber I963), seed
abortion has been attributed to other, largely
unknown factors (Dombrosky & Schowalter
I988, Schowalter et al. 1985). Unexplained
abortion typically ranges from 30-80% of
potential seed among Douglas-fir seed orchards
(Dombrosky and Schowalter I988, Schowalter et
al. 1985). A 50% increase in aborted seed
resulting from seed bug feeding, at densi- ties
comparable with those observed in seed
orchards, suggests that effects of L. occidentalis on
Douglas-fir seed production has been greatly
underestimated. (Schowalter and Sexton 1990, p.
1486)

Theisen (1976, p. 2) noted that damage at an


Oregon seed orchard had been severe, with
damage to full- size seed as high as 80%. It is
interesting that the only post-1990 reports
concerning insects and Douglas-fir seed known
to the present authors discuss the seed
destruction by Leptoglossus occidentalis.
Megastigmus spermotrophus
The genus Megastigmus in the family
Chapter 9. Cone and Seed Insects and Diseases 211
scale into which her ovipostor is inserted.
Hanson 1952, Keen 1958). The Douglas-fir seed
chal- cid (Megastigmus spermotrophus Wachtl)
probably has been the subject of more research
than any of the other insects known to damage
Douglas-fir seed. Undoubtedly, its life cycle,
which favors distribu- tion of the insect with the
seed, is the major reason for the relatively wide-
spread interest in this wasp. The geographical
distribution of this species fol- lows well the
distribution of its host. Infestation is reported on
Douglas-fir from California, Idaho, Colorado,
Washington, British Columbia, Oregon, and New
Mexico (Keen 1958). It was introduced in
infested seeds to Great Britain, western Europe,
and New Zealand. Jarry et al. (1997) discussed the
movement of Megastigmus in France, and its
inva- sion in seed orchards. This insect almost
certainly
feeds on seeds of Douglas-fir only.
The life history of Megastigmus is well
studied, but little is known about the ecological
factors affect- ing the distribution and intensity of
attack. Hussey (1955, 1956) published a detailed
study on the life history and habits of this species.
Mating takes place on the Douglas-fir needles
soon after the adults emerge from the pupal stage.
The female can lay fertile eggs without
fertilization, but all the adults from such
parthenogenetic eggs are males. The act of
oviposition was described by Miller (1916). He
reported that the female rests on a cone scale with
her head pointed toward the base of cone, drives
her ovipostor through the cone scales and deposits
an egg in a young seed. Two to five minutes are
re- quired for oviposition. According to Hussey
(1954), normally only one egg is laid in a seed,
but where there is considerable competition
between the egg- laying females for seed, as many
as seven eggs are found in one seed. Only one
larva develops to the adult stage when several
eggs are laid within the same seed. No specific
information is available on whether or not the
female would lay an egg into an unfertile or
empty seed, although it is known that a potentially
sound seed is necessary for the develop- ment of a
Megastigmus larva. It is conceivable that the
female selects by some means the fertilized seeds
for oviposition, which is indicated by the fact that
she spends a considerable time “choosing” the
period when the cones are subject to attack is
The earliest reports describing Megastigmus relatively short, either an unusually warm or
sper- motrophus Wachtl appeared in European cold spring may result in the peak of insect
journals. As previously noted, the insect was emergence and activity before or after,
first described and classified by Wachtl in
Vienna in 1893. Contrary to the general belief
of entomologists of the time concerning
Megastigmus spp., Wachtl noted that this
wasp is phytophagous. His conclusion was
sup- ported by a number of papers published
in the early part of the twentieth century
(MacDougall 1906 a,b; Crosby 1909, 1913;
Rohwer 1913; Miller 1914, 1916). These
reports outlined briefly the life history of this
insect as follows. The adults emerge in the
spring about the time that the young Douglas-
fir cones are turning down and are still
relatively unlignified. The female inserts her
ovipostor directly into the developing seed to
lay a single egg. The insect is apparently able
to detect the presence of the seed, but not
whether it contains an egg laid by another
Megastigmus. If more than one egg is laid in a
seed, only one larva will survive. The larva
develops during the summer, feeding on the
contents of the seed until, in the fall, it
occupies the entire seed cavity. The larva
overwinters in the seed. If the seed remains in
the cone, the larva is subject to parasit- ism by
the larvae of Amblymerus apicalis. Or, if the
seed is released from the cone, the larva is
subject to predation by seed eating mammals
(Hussey 1955). If the larva survives the
winter, it may pupate and the adult emerge in
a rather restricted period of 2 weeks (between
mid-May and mid-June in Britain), depending
upon the latitude (Hussey 1955, 1956).
Hussey noted further that the
development of Megastigmus is much more
closely controlled by temperature than is the
development of Douglas-fir cones. For
example, the time of earliest Megastigmus
emergence varied from May 20 to June 4 in
three consecutive years. But, during the
same years, the time of maximum
susceptibility to attack of the cones varied
only 5 days (June 8–13). It was found that
pupal development had a threshold tempera-
ture of 5.8°C and that about 390 degree-
days (in excess of 5.8°C) were required to
complete pupa- tion. Therefore, since the
212 Douglas-fir: The Genus Pseudotsuga
respectively, the optimum time for oviposition. sition all can affect the level of insect infestation.
Such years obviously provide a check on Findings in subsequent studies varied greatly, both
Megastigmus populations. However, low within a given species and between species, in level of
temperatures during the spring months frequently insect predation upon Douglas-fir seed. Kozak
increase the number of larvae that remain in
diapause for a second or third year, a strategy
which appears to be the insect’s mo- dus operandi
for dealing with the widely fluctuating production
of cones by Douglas-fir from one year to the next.
And, not only does cold weather delay the
development of the adult insect, it inhibits the
ability of the female to oviposit, according to
Miller (1964), who found that the females were
active primarily during sunny days in southern
Oregon.
Hussey (1955, 1961) observed that the
optimum time for oviposition differed between
the cones of variety menziesii and variety glauca
trees in Britain. The former were successfully
attacked after they were pendant, 4 to 8 cm in
length, and until all but the apical fringe of the
scales was brownish. This period commonly
extends from 2 to 3 weeks. The cones of the
variety glauca trees are most susceptible to attack
when the cones are nearly full grown, 4 to 6 cm in
length, and the ovuliferous scales are a deep
mauve pink.
The larvae hatch a few days after the eggs
are laid. Larval development proceeds through
five instars until, at the end of 6 to 7 weeks, the
mature larva has eaten the contents of the seed
and occupies the entire megagametophyte
cavity. The seed coat and testa develop
normally, and the infested seed cannot be
separated from sound seed by external
examination. Larvae then enter a resting stage,
which most commonly lasts until the following
spring; in cooler climates such as northern
Britain, however, as many as half of the larvae
may not pupate for an additional year or two.
The delayed pupation means that the level of
infestation for any one year is related to the
levels of infestation and the size of the cone
crop for the previous two years. Roux et al.
(1997) noted that low temperatures described as
“chilling” are required for diapause to be
complete (p. 176).
Graham and Prebble (1940, 1941) noted that
crown position, cone size, tree age, and tree po-
(1963) reported a significant level of damage occasioned by M. spermotrophus is a function not
by M. spermotrophus in cones collected in only of the frequency of such attacks, but also of
coastal British Columbia and no evidence of the degree of fertilization of the seeds.
damage by B. col- faxiana. In contrast, Hedlin Accordingly, previous estimates of insect damage
(1964a) working in the same area, reported have been ei-
substantial damage by B. col- faxiana and only
light predation by M. spermotrophus.
Schowalter et al. (1985) noted increasing
damage by all insects from north to south
between British Columbia and northern
California. And Baron (1971,
p. 491), noted heavy insect damage of
Douglas-fir in California. Finally, both Hedlin
and Ruth (1978) and Schowalter and Haverty
(1989) reported clonal differ- ences in Douglas-
fir in resistance to M. spermotrophus attack.
However, the former authors noted that such
differences probably have no practical
significance. Hofmann (1924) observed that
“when the seed crop is light the seeds are
generally attacked by an insect (Megastigmus
spermatotrophus Wachtl), which destroys
before maturity a large percent of the few
seeds which would otherwise be produced”
(p. 49). Overall the degree of seed damage by
M. spermot- rophus reported varies from less
than 10% of sound seeds to more than 50% in
western North America to almost 100% in
Europe (Jarry et al. 1997, Krístek 1967,
Lessman 1974). The much higher figure in
Europe may reflect the fact that M.
spermotrophus has no competitors in Europe,
whereas it does not compete well in North
America against B. colfaxi- ana (Volney
1984) or C. oregonensis (Rappaport and
Volney 1989).
Later reports (Niwa and Overhulser 1992,
Rappaport et al. 1993, Skrzypczynska 1994),
how- ever, suggested that the earlier reports,
which based measures of damage by M.
spermotrophus on the hypothesis that the
insect attacked only those seeds whose female
gamete had been fertilized, were in- correct.
They found that when an egg of M. spermot-
rophus is laid in a seed that contains an
unfertilized gamete, the seed will continue to
develop. Normally, the contents of seeds of
Douglas-fir that are not fertilized will be
resorbed by the plant (Allen and Owens 1972)
and such seeds will then be tallied as empty.
This suggests, then, that the level of damage
Chapter 9. Cone and Seed Insects and Diseases 213
In a 6-year study Hedlin (1964a) found that the gall
ther slightly high, when there was a high level of midge larvae were always more numerous than those
successful fertilization, to as much as 50% too of other insects.
high, when there were low levels of pollen.

Contarinia oregonensis
Hedlin (1974) and Miller and Ruth (1989)
reported that the Douglas-fir gall midge,
Contarinia orego- nensis, is one of the most
serious pests of Douglas- fir cones, particularly in
the wetter portions of its range in British
Columbia. Owston and Stein (1974) agreed that it
is a serious pest of Douglas-fir seed. Schowalter et
al. (1985) noted an increasing incidence in western
Oregon and Washington, however, and reported
that C. oregonensis and M. spermotrophus
together had destroyed 70% of the filled seed in
seed orchards (1985, p. 1228). As many as 30 C.
oregonensis larvae may form a single gall, which
will destroy both seeds on the scale. More
commonly, the major damage is caused by the
galls, which make extrac- tion of the seeds from
the cone impossible.
The adult insects emerge from cocoons in the
litter
in early spring and are the first insects to lay eggs,
frequently before the abortion of conelets is
complete. Volney (1984) reported that conelet
abortion was 50% in 1980 and 30% in 1981 in
California. The midge lays eggs near the ovules.
The eggs hatch in 2 to 3 weeks, when the larvae
invade the seed tissue and cause a gall to form
around the seed. This gall may either fuse the seed
and cone scale, which either effectively prevents
the extraction of the seed or inhibits further seed
development. Johnson (1963a) noted that the cone
midge is capable of attacking the cones only while
they are receptive to pollination, a period of 7 to
10 days. As a consequence, a major portion of the
mortality of the larvae may be caused by death of
the cone before the development of the insects.
Although the cone moth, B. colfaxiana is
capable
of destroying the gall midge galls and larvae, the
two species apparently can co-exist in the same
cone. However, Stein et al. (1988) reported data
suggesting Lepidoptera predation of Contarinia, and
Miller (1984a) noted that numbers of midge
larvae were reduced from 193 to 8.3 per cone in
cones with 2 Lepidopterous larvae per cone. This
may occur only when there are light cone crops.
when the cold mist caused a greater than 10-day
One possible reason is that Contarinia is delay in bud activity) may have been due to the
capable of colonizing the sterile cone scales at fact that the midge was able to oviposit during the
the base and tip of the cone, which are not application of mist (Miller 1983).
attractive to its com- petitors (Volney 1984).
Even so, however, several studies (Rappaport
and Volney 1986, Miller 1986a, Schowalter
and Sexton 1990) reported a high propo- tion
of C. oregonensis larvae in the mid-cone area.
Once the gall is formed, the larvae remain
in a V-shaped configuration, where they
obtain most of their food by absorption
(Hedlin 1961a) until the fall, when they
complete their development and it drops to the
ground. There are three larvae instars (Hedlin
1961a). The larvae will not leave the cone
until fall rains moisten it and low
temperatures favor the exit of the larvae
(Hedlin 1959). Often, larvae will select an old
male cone in which to make a cocoon and
overwinter; however, C. oregonensis may
have a diapause as long as 4 years (Miller and
Hedlin 1984, Danks 1987). Both Miller
(1986a) and Schowalter et al. (1986) reported
clonal differences in the infestation of
Douglas-fir cones by the cone gall midge.
But, while the former noted that such
differences were not significant, Schowalter et
al. (1986) reported statistical significance in
their data and suggested that the two-fold
difference between families with high
infestation and those with low is heritable.
About half of the pupae enter diapause each year
(Johnson and Winjum 1960, Hedlin 1961a,
Hedlin et al. 1980, Schowalter 1984). This
trait, together with the fact that the insect is
well adapted to its environment, assures
continuing populations, as Hedlin (1961a)
observed: “adults are active over a wide range
of temperatures and so are able to deposit
eggs even under relatively unfavorable
weather conditions. In the autumn, larvae are
able to survive for long periods in dry cones,
and after leaving the cones, in winter. Larvae
concentrate in litter below the tree. This
concentration of population probably
facilitates mating when males and females
emerge in the spring” (p. 965). Given the
foregoing, the lack of consistent effect of cold
misting upon midge in- festations (levels of
insect attack were significantly reduced only
214 Douglas-fir: The Genus Pseudotsuga
Two parasites of Contarinia oregonensis, Contarinia washingtonensis. He noted that C.
Torymus sp. and Platygaste sp., have been washingtonensis is generally smaller than C.
observed, but neither caused a high level of oregonensis, that the
mortality (Hedlin 1961). These data were
confirmed by Miller (1984b), who noted that
numbers of eggs deposited accounted for 73% to
100% of the variation in midge populations, and
that the incidence of parasitoids was small.
However, Johnson and Heikkenen (1958) noted
that an un- identified species (Chalcidoidea)
parasitized 40% of
C. oregonensis in some cones. Miller also
suggested that predation by the larvae of cone
moths or worms may have accounted for some
midge mortality, but that neither are commonly
present in high numbers on the British Columbia
coast.

Contarinia washingtonensis
The presence of a second species of Itonididae in
Douglas-fir cones was first reported by Johnson
and Heikkenen in 1958, but the insect was not
identified. The following year, Hedlin (1959)
described the Douglas-fir cone scale midge,
Contarinia washingto- nensis Johnson (Diptera:
Cecidomyiidae), as follows: “Although this midge
lives in close proximity to Contarinia
oregonensis in Douglas-fir cone scales, it can be
readily separated on the basis of appearance and
habits” (p. 10). He noted that larvae of this insect
were very plentiful and that mature larvae were
deep orange in color, compared with the quite pale
color of C. oregonensis. When feeding, the larvae
lies along the cone scale fibers and does not form
a gall. “Damage was observed readily in July
when the larvae were reaching maturity. The areas
where damage occurred turned brown, and
expanded as feeding increased. By the time the
larvae are fully grown, they may be lying almost
fully exposed on the inner surface of the scale,
and at this time leave the cone readily”; and that
there were as many as 36 larvae found per scale.
“Examinations were not carried out to determine
the extent of damage caused by these insects but
there is no doubt that they contribute to seed loss”
(Hedlin 1959, p. 11). However, Miller and Ruth
(1989, p. 30) found that
C. washingtonensis rarely consumes seed.
Subsequently, Johnson (1963a) published a de-
scription of this insect, which he named
adults lay eggs beneath the long bracts after small clusters in the angle between the cone bract
the cones are pendant and closed (about 3 to 5 and scale. Young larvae tunnel into the cone scale
weeks later than C. oregonensis), and that the tissue where they feed in the central part of the
larvae leave the cones early in the fall, in scale, usually in small
contrast to those of C. oregonensis, which
remain in the cones until onset of fall rains. A
further difference between the species is that
C. washingtonensis larvae feed on the cone
scales at some distance from the seeds and do
not form galls. Hedlin and Johnson (1963)
confirmed the foregoing, noted that the
orange colored larvae of
C. washingtonensis seed fully extended rather
than in a curved position, and reported
damage by this species in western
Washington of up to 47% of the seed. They
suggested that this insect has a more flexible
life cycle than that of C. oregonensis and may
pose a significant threat to seed orchards.
Hedlin and Johnson (1963) contended that
“the midge, Contarinia washingtonensis is
capable of causing serious seed loss in
Douglas-fir” (p. 1168). Johnson and Winjum
(1960) described the Douglas- fir scale midge
as “an undescribed species found at- tacking
the cones of Douglas-fir. It has been abundant
the last two years in Washington and also in
British Columbia. The adults emerge from the
soil where lar- vae have over wintered in
cocoons and deposit their eggs under the bract
of the pendant cone in early June. The larvae
upon hatching from the egg bores into the
cone scale where they feed” (p. 10). Johnson
(1963a) presented a detailed anatomical study
of this insect, described the species, and
reported the morphological characteristics of
the different stages (see also Hedlin and
Johnson 1963). The females lay eggs when
the cones are closed and pendent. The eggs
are laid beneath the bracts of cones. The
newly hatched larvae mine in the cone scales
and do not cause galls, like Contarinia
oregonensis larvae, when the cones are still
greenish. The authors could find no
information on either the ecology of the
insect, or on the factors affecting the insect in
the different stages of development.
According to Johnson and Hedlin (1967), “adults
emerge during late May and early June to lay
their eggs in young cones about the time
foliage buds are bursting. The eggs are laid in
Chapter 9. Cone and Seed Insects and Diseases 215
remains dormant over winter to resume
groups. Infested scales turn brown before the
cones are mature, but galls are not formed” (p. 5).
Theisen (1976) reported, however, that the
“Contarinia midge adult lays eggs near ovules
when flowers are open for pollination. Damage to
seed slight, usually no external evidence of
damage” (p. 4).

Dioryctria abietella
Yates (1984) gave the following overview of
Dioryctria abietella Denis & Schiffermüller
(Lepidoptera: Pyralidae, Phycitinae):
If there is one genus universally identified as a conifer
cone and seed destroyer, it is Dioryctria—the cone
worm. Best known is the cone pyralid, D. abietella
(Denis and Schiffermüller), which is one of the most
widespread cone and seed insects in the world. The
distribution of
D. abietella includes Europe westward through Asia
to Japan and southward into Pakistan and India. Hosts
include nearly all conifers growing within this region.
Because of this species’ wide distribution and varied
host preferences, the biology varies considerably.
(Yates 1984, p. 33)

Dioryctria abietella has not been intensively studied


on Douglas-fir. The work that has been done indi-
cates the wide range of distribution and habits of
the insect. The wide distribution of the species is
shown by the fact that specimens in the United
National Museum are from Abies spp., Douglas-
fir, and all Pinus spp. in North and Central
America, from the Canadian provinces of British
Columbia and Newfoundland and Labrador south
to Guatemala. Lyons (1957) noted that D. abietella
“infests the cones, shoots, and bark of many
conifers, and is apparently Holarctic in
distribution” (p. 71). It may produce “two
generations per year in the western United States”
(p. 76), and has been reported over- wintering in
several different stages, partly grown larvae,
prepupae and pupae. According to Johnson
and Hedlin (1967),
This insect occurs sporadically but is capable of
causing considerable destruction. The moth usually
emerges in spring but may emerge in the fall. The egg
laying habits are not fully known, but eggs laid in the
spring hatch in early summer. The larva feeds in a
manner similar to that of the Douglas-fir cone moth,
except that it feeds throughout the cone and one larva
may destroy a cone completely. Large quantities of
frass are common on the surface of infested cones
(Figure 2). The larva is larger and darker in color than
that of the Douglas-fir cone moth. In the fall, the
mature larva leaves the cone to spin a soft round
cocoon in which to overwinter. The immature larva
for a wide range of climatic condi- tions. Because of
feeding in the spring and complete its its variable feeding habits, population levels of
metamorphosis in late summer. (Johnson and Dioryctria abietella are probably relatively inde-
Hedlin 1967, p. 2) pendent of the supply of cones and thus are not likely
to suffer from competition for food supplies.
The life history of Dioryctria is not yet
clear. Keen (1958) also noted that there may
be two generations per year. With global
warming there will prob- ably be two
generations in the range of Douglas-fir. Lyons
(1957) found only one generation per year in
Ontario. As Keen (1958) described, some
eggs are deposited by moths reaching the
adult stage in October. More eggs are
deposited by another group of moths that
emerge in May or early June. After hatching,
the larvae bore through the scales and feed
indiscriminately on scales, bracts, and seeds.
As the cones ripen, the larvae leave them and
form their cocoons on the ground. Some of
them pupate immediately and emerge in
October; the rest of the population (probably
most of them) spends the winter as prepupal
larvae, and pupates and emerges the following
spring. No information is available on
diapause in this species.
The eggs are 1 mm long, white, oval, and flat-
tened with finely roughened surface. Larvae
have five instars with respective headwidths of
0.45, 0.71, 1.10, 1.35, and 1.70 mm (Lyons
1957). They are red or purple in color,
sometimes with a greenish tinge. The absence
of anal comb helps to distinguish Dioryctria
abietella from the Douglas-fir cone phalo- niid
(Henricus fuscodorsana (Kearf.). The pupa is
10 to 12 mm long, with straight, slender caudal
hooks. The forewing of adults is predominantly
grey, with white transverse zig-zag lines. The
wing spread of adults was described as 23 to 28
mm by Lyons (1957). Koerber (1960) stated that
“the larvae of Dioryctria abietella usually do not
bore as deeply into the cones as do the larvae of
Barbara colfaxiana. As a result less seed is
destroyed by Dioryctria abietella” (p. 11). But
Johnson and Hedlin (1967) noted that the larvae
feed throughout the cone and that one may
destroy all
of the seed.
According to Koerber (1960),
There is almost no information on ecological
factors which might affect this species. Only six
species of para- sites and predators are known to
attack it in western North America. Its wide
distribution might be taken to indicate a tolerance
216 Douglas-fir: The Genus Pseudotsuga
Kozak (1963) noted that the level of damage to damage caused
Douglas-fir cones by D. abietella was highly
variable (5.7% to 18.2% cones damaged) and that
this species was less important in coastal British
Columbia than were Contarina spp. or
Megastigmus spermotrophus. Hedlin (1958) did not
include it in his survey of in- sect damage in
British Columbia; Schowalter et al. (1985) noted
that D. abietella was not a major pest of Douglas-
fir seed and cones in western Washington and
Oregon, but that the damage from this insect
increased from north to south. Ruth (1980) found
that while rarely present in large numbers, larvae
of D. abietella may destroy all the seeds in a cone
by indiscriminate feeding on cone scales and
seeds. The damage is characterized by large holes,
coarse frass on the exterior of the cone, and, in the
case of Douglas-fir, almost complete destruction
of the cone (Hedlin 1974).
Fatzinger and Asher (1971) reported that D.
abietella males were attracted to females by a sex
pheromone; Coulson and Franklin (1970)
presented a review of a closely related species, D.
amatella, and emphasized the complexity of the
life cycle and the great variety of structures
colonized. Although Rappaport and Volney
(1989, p. 146) suggested that damage by C.
oregonensis may be reduced by D. abi- etella,
their data showed that any such differences were
not significant.

Lepesoma lecontei
Schowalter (1986) and Dombrosky and
Schowalter (1988) suggested that a flightless
weevil, Lepesoma lecontei (Coleoptera:
Curculionidae), may have been responsible for
seed losses (5.8%) equivalent to those caused by
Megastigmus or Contarinia in seed orchards of the
Pacific Northwest in 1984 and 1985. Surveys of
arthropods emerging from Douglas-fir litter in a
50-year-old Douglas-fir stand in seed orchards
(Schowalter 1984a) demonstrated that populations
of Lepesoma were greater in seed orchards, so it is
not clear (1) whether the seed orchard data
reflected conditions in “wild” forests and (2)
whether the management of the seed orchards in
some way fa- vored the life cycle of Lepesoma. It
is interesting that the damage to the early stages of
ovulate strobili strongly resembled frost damage;
this may have been the reason that the extent of
by this insect was not recognized earlier. Diapause
Schowalter (1986) demonstrated that the
damage caused by L. lecontei was the result Dormancy in insects or “diapause” has been
of conelet destruction rather than feeding on described as a ”’physiologically’ controlled
a seed. programmed rest

Choristoneura occidentalis
The western spruce budworm, Choristoneura
oc- cidentalis Freeman (Lepidoptera:
Tortricidae), is commonly considered to be a
defoliator rather than a predator of cones and
seeds. But studies from the Rocky Mountains
have reported that this insect can be
responsible for significant destruction of
develop- ing cones. The first report of
budworm damage to Douglas-fir cones
(Dewey 1970) indicated that this insect
destroyed an average of 36% of the seeds in
Montana and in Yellowstone National Park
and that it was by far the most numerous of 13
seed and cone insects collected. He noted that
several second- instar larvae might be found
in one cone with other insects. But a single
fifth- or sixth-instar larvae could destroy all
the seeds in a cone and might consume the
larvae of Megastigmus or Contarinia. Larval
de- velopment was generally complete by
mid-July, when the larvae formed a pupae,
sometimes in the hollowed-out remainder of
the cone. Shearer (1984) confirmed Dewey’s
(1970) data for western Montana, although the
proportion of Douglas-fir seed reported
destroyed by budworm was lower. Finally,
Frank and Jenkins (1987) found that the
western spruce budworm caused significant
damage to all repro- ductive structures on
Douglas-fir in west-central Idaho, and that the
number of seeds destroyed was exponentially
related to the degree of defoliation this insect
caused. According to Reardon et al. (1985),
In the northern Rockies, the most serious pest
affect- ing Douglas-fir cones is the western
spruce budworm, Choristoneura occidentalis.
Early instars of the western spruce budworm
can destroy conelets before and after
pollination, whereas later instars feed on the
many developed cones. Larvae often feed on
more than
one cone during their development, and most
seeds in infested cones are either destroyed
directly or retained within cones because of
growth distortions and excess rosin exudation
induced by insect feeding. (Reardon et al. 1985,
p. 961)
Chapter 9. Cone and Seed Insects and Diseases 217
3. The term “diapause” was not in use at that time.
of growth, development, or reproduction resulting
in a reduction of active physiological functions”
(Danks 1987, p. 8). It is discussed in detail in
Danks (1987); however, the following points
drawn from that reference are helpful for this
discussion (num- bering added):
1. Insect life cycles are timed so that active stages
coincide with favorable conditions so that periods that
do not provide requirements for development can be
passed safely. (Danks 1987, p. 4)

2. Diapause affects food storage, largely fat and general


biochemistry. (Danks 1987, p. 19)

3. Insect structure is more resistant to drought in


dia- pause. (Danks 1987, p. 23)

4. [D]iapause has coincided times with cold


hardiness but is not interdependent. (Danks 1987, p.
41)

5. Diapause is primarily cued by photoperiod


although there may be interactions with
temperatures, moisture and other environmental
parameters, light intensity required generally less
than 1 lumen, affected by light quality. (Danks
1987, p. 230)

Danks (1987) emphasized the great varieties of


terms associated with diapause and the difficulties
engendered in research efforts designed to
establish parameters. Stadnitskii (1986)
concluded,
Hence, we may consider the conobiont diapause as an
adaptation to cone crop dynamics, appointed by
natural selection. Carpobionts of leaf-bearing tree and
shrub spe- cies, as a rule, have no diapause, excluding
C. glandium and some other strictly adapted species.
Cone insect diapause and tree reproduction
dynam- ics are induced by similar natural factors. We
may con- sider them to be a total result of a long-term
coevolution of coniferous species and their
phytophagous insects. (Stadnitskii 1986, p. 244)

Perhaps the first mention of dormancy3 in


western cone and seed insects was that of Miller
(1914), who noted that as much as half of an
insect brood may be dormant for 2 or more years.
In a study focused on Megastigmus strobilobius
and the cone moth, Cydia strobilella, in Finland,
Annila (1984) observed that “a high proportion of
cone and seed insects do not emerge after one
hibernation but remain in diapause for several
years,” and that the “duration of diapause is an
important factor having an effect on population
fluctuations of cone and seed insects. Prolonged
diapause has been considered to be a means of
development the summer prior to year of maturation
adapting to the varying cone crop of the host also affect inversely the induction of prolonged
tree” (p. 57). Annila concluded that “it seems diapause. This relation may be necessary for the
survival of B. colfaxiana populations, in that prolonged
possible that seed insects have not only diapause prevents excessive intraspecific competition
adapted to fluctuations in the annual cone for food resources in years of low food abundance.
(Hedlin et al. 1982, p. 468)
yield but also the fluctuations in the cone
destroyer population as well” (p. 63). Hedlin
(1964a), working with cones infested with
insects on Vancouver Island, noted that “apart
from availability of cones, the phenomenon of
diapause was probably the most important
single factor in- fluencing insect population
fluctuation” (p. 124). In examining diapause
in Barbara colfaxiana, Sahota et al. (1982)
reported the following:
Pharate adult (adult within the pupal cuticle)
diapause was discovered in Barbara colfaxiana.
This phenomenon is uncommon in insect
development. It was also shown that both the
termination of diapause and the subsequent
advancement of pharate adult development can
occur at 0°C. In many insects, the normal
progression of the
life history is interrupted by a state of dormancy
resulting in a discontinuation of growth
development. The most advanced of these
dormancies, induced well before the onset of
adverse conditions and maintained for sometime
irrespective of environmental conditions, is
commonly known as diapause. (Sahota et al.
1982, p. 1179)

According to Hedlin et al. (1982),


Studies were carried out in the field and in the
laboratory to determine if prolonged diapause of
the Douglas-fir cone moth, Barbara colfaxiana
(Kearfott), was correlated with the sizes of the
cone crops maturing in the year of larval feeding
(N) and in the year following larval feeding
(N+1), and to determine if weather during the
period of larval feeding (year N) influenced the
size of the maturing cone crop or the incidence of
prolonged diapause the fol- lowing year (N+1).
Field studies showed that prolonged diapause
induction in B. colfaxiana was not rank corre-
lated (but approached significance) with the size
of the cone crop maturing in year N, but was
negatively rank correlated with that in year N+1.
Two of seven weather parameters, mean
maximum temperature and mean daily
temperature, measured during the larval feeding
period were positively rank correlated with cone
crop size. No parameter was correlated with the
incidence of prolonged diapause. In the
laboratory, the incidence of prolonged diapause
was negatively correlated with temperature.
Photoperiod and parental diapause habit had no
direct effect. (Hedlin et al. 1982, p. 465)

The relationship between prolonged diapause and


cone crop size the year following larval seeding
suggests that factors which affect cone
218 Douglas-fir: The Genus Pseudotsuga
They noted that both Barbara and Contarinia autumn and winter, but these temperatures are not
larvae also frequently go into diapause. given, nor is their dura- tion. Miller et al. (1984, p.
In other reports concerning diapause on 49) found no apparent
Barbara colfaxiana, Sahota et al. (1983) noted
that the pharate phase of insect development starts
within 24 hours after pupation in both 1-year and
2-year diapause individuals. Sahota et al. (1985)
found that year-1 and year-2 diapause of B.
colfaxiana can be identified by the color of
individuals and anatomical features. Sahota and
Ibaraki (1991) showed that diapause was not
related to insect dry weight, but that 2-year
diapause insects had a slightly higher dry-weight
to fresh-weight ratio.
Radcliffe (1952) reported that in the Cowichan
Lake area, from 3% to 11% of the pupa of B.
colfaxiana were in diapause. Miller and Hedlin
(1984) noted that “induction of prolonged
diapause in the cone moth in a dry interior area
was correlated inversely with cone crop size the
year following larval feeding and directly with
cone crop size the years of larval feeding” (p. 91).
Miller and Ruth (1986) noted that “high
temperatures during May to August terminate
prolonged diapause” of B. colfaxiana (p. 1073),
and that “the termination of prolonged diapause
by summer temperatures would allow for greater
moth emergence when larger cone crops are
produced as production of cones by Douglas-fir is
also positively correlated with summer
temperature the year prior to cone maturation” (p.
1074).
Roux et al. (1997) reviewed the literature
concern-
ing diapause, particularly in spermatophases,
dis- cussed the relation of temperature and
photoperiod to simple and prolonged diapause
in Megastigmus spermotrophus. They concluded
that “in the Douglas- fir seed chalcid, chilling is a
prerequisite for the completion of the
development of the two kinds of diapause, low
temperatures in autumn and win- ter activating
the development of both simple and prolonged
diapause. This finding indicates that
environmental information can influence the
nature of larval diapause (simple or prolonged),
but we cannot eliminate other factors, such as
genetic fac- tor (genetic polymorphism) or ‘bet-
hedging’ strate- gies” (p. 176). The “chilling” in
this work refers to natural temperatures in
correlation between prolonged diapause and The first attempt to control insects that feed on
cone crop size for Douglas-fir cone gall midge Douglas-fir seeds was that of Rudinsky, who
at Lake Cowichen. In an early report of applied
diapause in cone and seed insects, Hussey
(1956, p. 193) noted that 68% and 76% of
Megastigmus larvae in England extended
diapause in 1952 and 1954, respectively, but
only 6% did so in 1951.

Control
For many years, when seeds were collected
from wild stands, control of seed and cone
insects was not a problem. Although it was
generally recognized that insects could
destroy significant quantities of Douglas-fir
seed, efforts to control such damage began
only in the 1950s, with the advent of costly
seed orchards in the northwestern United
States (Hedlin 1961b, Schowalter et al. 1985).
Surveys
With tree breeding programs came increasing
em- phasis on protecting valuable seed crops,
as well as more detailed evaluation of
potential damage to aid in determining
whether preventive treat- ment was justified.
A number of papers proposed evaluation
techniques utilizing the principals of
sequential surveys to estimate parameters
such as cone and seed efficiency. Such studies
identified seed losses caused by insects
previously not im- plicated as seed predators,
established the relative importance of known
and unknown environmental factors and
insects in reducing potential seed crops,
submitted cost analyses of seed production,
and provided a measure of the effectiveness of
chemical treatments (Miller 1983, Dombroski
and Schowalter 1988, Schowalter and Sexton
1989). Research in this area has included the
development of a sequential sampling system
of seeds to estimate predation by Contarinia
oregonensis Foote (Miller 1986b) and
Megastigmus spermotrophus (Kozak 1964),
the dis- tribution of the eggs of Contarinia
oregonensis and Barbara colfaxiana (Miller
1986c, Sweeney and Miller 1989), and the
development of a partial life table for Barbara
colfaxiana (Nebeker 1977 and Miller 1989).
Artificial control
Systemic and chemical
Chapter 9. Cone and Seed Insects and Diseases 219
Nonetheless, the results were erratic and varied with
DDT in a water emulsion with ground-based chemical and insect: dimethoate was not effective in
spray equipment in 1955 (Koerber 1960). The any study; acephate
results were inconclusive, but there appeared to be
a reduction in the damage caused by Barbara and
Megastigmus, but not by Contarinia. In the years
that followed, a number of reports (Johnson
1963b,c; Koerber 1963; Hedlin 1964b; Johnson
1964; Johnson and Rediske 1964, 1965; Buffam
and Johnson 1966; Hedlin 1966; Johnson and
Meso 1966; Johnson and Hedlin 1967; Johnson
and Zingg 1967) were published that de- tailed
efforts to employ either ground-based spray
apparatus or aerial systems to treat Douglas-fir
trees with a range of insecticides. Chemicals used
included contact insecticides such as lindane and
dieldrin and systemics such as Meta-Systox R,
Bidrin, and Dimethoate. While investigators
reported that in most of the trials, the chemicals
appeared to be ef- fective in reducing seed loss, as
Miller (1980) noted, the results were erratic,
affected by weather, varied with chemical used
and insect targeted, weakened by the fact that they
rarely presented statistical analy- ses, and suffered
from the lack of efficient sampling designs to
estimate insect populations.
The past 30 years have seen the emphasis in
research to control insect seed predators shift
from aerial or ground applications of insecticides
to the development of (a) techniques to inject
systemic in- secticides into tree boles; (b)
sampling methodology to provide accurate
measures of the need for insect control prior to
implementing control measures; and
(c) alternative systems to manage insect
populations. The shift toward the injection of
insecticides was occasioned by the fact that such
an approach was not subject to a specific time to
be effective and, because the chemicals were
enclosed in capsules, the danger to the applicator
was reduced over that associated with sprays.
Several papers reported the effect of injected
systemic insecticides upon target insect
populations (Schenk et al. 1967, Johnson et al.
1984, Koerber and Markin 1984, Reardon and
Barrett 1984, Reardon et al. 1985, Stein and
Markin 1986, Stein et al. 1988, Stein and Koerber
1989, Stein et al. 1993, de Groot et al. 1994).
These trials were superior to the earlier work in
that they all had com- petent statistical designs.
trees. Miller (1983) noted that when the delay in
demonstrated positive control of only reproductive bud break was at least 10 days and
Dioryctria; and oxydemeton-methyl when early bud break trees were the primary
(Metasystox-R) was generally effective against targets of Contarinia oregonensis Foote,
Barbara, Dioryctria and Contarinia, but not
Megastigmus. It is possible that the negative
results with Megastigmus reflected the fact
that this insect competes poorly with other
seed predators (Rappaport and Volney 1986,
1989; Volney 1984). Miller (1986a) noted that
chemicals were the only practical method for
controlling insects in Douglas- fir seed
orchards at that time. Schowalter (1984b,
p. 1437) reported that both C. oregonensis and
M. spermotrophus were capable of dispersing
over at least 85 m, and that destruction of
debris within seed orchards would not protect
them against these insects. Summers and Ruth
(1987) found that sprays of permethrin and
dimethoate, but not diatomaceous earth, were
effective against L. occidentalis.
Chemical attractants
Although chemical attractants (pheromones)
have been used in the management of
mountain bark beetle (British Columbia
Ministry of Forests 1995a) and defoliators and
nursery pests (Lavender et al. 1991), there is
little evidence of their use in manage- ment of
pests of Douglas-fir cones. Hedlin and Ruth
(1968) demonstrated that male Barbara
colfaxiana moths were attracted by females,
and Hedlin et al. (1983) showed that mixtures
of (Z)-9-dodecyl acetate and dodecanol were
effective in attracting male Barbara moths.
There were no published reports on the use of
these materials in the management of seed
orchards, however.

Phenology
Cold water spray
Trials by Silen and Keane (1969)
demonstrated that spraying Douglas-fir seed
orchards with cold water during the period of
reproductive bud develop- ment could
delay such growth by as much as 12 days
compared with that of unsprayed trees.
These results stimulated later tests in British
Columbia designed to determine whether
such asynchrony might reduce the incidence
of predatory insects in cones of treated
220 Douglas-fir: The Genus Pseudotsuga
damage by this insect could be reduced as if any, on seed viability, although the later
effectively as by applications of the insecticide
dimethoate. However, the delay occasioned by the
spray treat- ment was affected by weather patterns
during the spray period—low temperatures
reduced the effect of the cooling treatment—and,
hence the effects of spraying upon levels of insect
damage could not be predicted. The author
suggested that accurate determinations of the heat
sum requirements of the insect and of the seed
orchard were needed. El- Kassaby et al. (1990)
reported significant reductions in the incidence of
Megastigmus spermotrophus Wachtl larva when a
Douglas-fir seed orchard received the cold water
spray.
Barriers
Zhang and Schowalter (1997) noted that physical
barriers such as sticky material (tanglefoot),
which reduced weevil damage to from 25% to
5%, may be effective in some locations. We know
of no other reports describing such control,
however.
Seed treatment
Hussey (1954) noted that the larva of
Megastigmus spermotrophus Wachtl could be killed
by the tempera- tures employed to dry cones prior
to seed extraction, but did not quantify the
treatment. Later trials by Ruth and Hedlin (1974)
examined the efficacy of a range of temperatures
and treatment times to kill Megastigmus larva
and determined that exposure of infested seeds to
a temperature of 45 C for 40 hours resulted in
100% mortality of the larva without sig- nificantly
affecting seed germinative vigor. They cautioned,
however, that seed moisture content (MC) should
not exceed 9% at time of treatment initiation.
Richardson and Roth (1968) demonstrated that
exposure of Megastigmus infested Picea abies (L)
Karst seed to hydrocyanic acid for 2 hours was
sufficient to kill the larvae, but that this insect
demonstrated significant resistance to methyl
bromide. Further trials (Roth and Strasser 1971)
confirmed the resis- tance to methyl bromide and
showed that a mixture of carbon disulfide and
carbon tetrachloride was effective against larvae
of both Megastigmus spp. and Cecidomyiidae in
seeds of Douglas-fir and Port Orford cedar,
respectively. Neither report discussed the effects,
paper noted that another researcher had found heavily parasitized by Contarinia spp.
no ill effects of carbon disulphide on Douglas- 3. Levels of Megastigmus are significantly higher in
fir seed germination. Finally, Sweeney et al. Europe despite the fact that parasities by autochtho-
(1991) reported that the incubation drying
separation method was effective in separating
most of the infested seeds from sound seeds,
in that the infested seeds floated after
treatment whereas the sound seed sank.

Natural control
Schowalter (1984a) noted that the destruction
of the litter where many insects overwinter
will not protect seed orchards, most of which
are within the range of insects bred in wild
stands. Schowalter et al. (1985) noted that the
insects discussed here, “frequently destroy
over half of the Douglas-fir seed crop” (p.
1223).
Competition
Because Douglas-fir cones are commonly
attacked by several insects, the question of the
extent to which the various species limit or
impact the damage each causes is of both
theoretical and practical inter- est. The
success of chemical control methods, for
example, depends on whether the elimination
of a target insect, such as Contarinia spp.,
which attacks the conelets early in the
growing season, may affect the damage
caused by Megastigmus spermotrophus,
which may deposit eggs after the insecticide
has been metabolized by the tree or otherwise
rendered innocuous. Also of interest is the
relative level of damage caused by
Megastigmus spermotrophus in western North
America, where it must compete with other
insects, and in Europe, where it has no
competition for Douglas-fir seeds.
For the following reasons, Rappaport and
Volney (1989) suggested that competition
between insect species indigenous to Douglas-
fir cones may affect the level of damage of
any one insect:
1. Competition occurs most often when the food
resource is limiting and the erratic occurrence of
Douglas-fir cone crops, together with the
extended diapause habit of all save Dioryctria
abietella, suggest that quantity of seed available
frequently limits insect populations.
2. Both Contarinia spp., and Megastigmus
spermatotro- phus have no defense against
predation by Dioryctria or Barbara, and
Megastigmus is generally not found in cones
Chapter 9. Cone and Seed Insects and Diseases 221
para- sitized by Glypta evetriae. Miller and Ruth
nus insects in Europe is higher than that occasioned
by the depauperate North American parasite
(1986)
complex and there are no known pathogens of
Megastigmus on either continent.

4. The tentative evidence from chemical control proj-


ects suggest that when one insect is controlled, levels
of competing insects rise to occupy the niche created.
(Rappaport and Volney 1989)

Given the foregoing, research in northern


California has shown that potential competition
between Barbara and Contarinia spp., was
reduced because the former feeds primarily in the
center of the cone, whereas the latter favors the
ends (Volney 1984)—although Rappaport and
Volney (1986), Schowalter and Sexton (1989),
and Miller (1986a) all reported a majority of
Contarinia larvae in the center of cones; however,
Barbara and Megastigmus larva may coexist in
the central portion of the cone (center between the
ends). Subsequent trials (Rappaport and Volney
1986) confirmed the hypothesis that intra- species
competition was minimized by the spatial pattern
of insect feeding within the cones; Rappaport and
Volney (1989) found, however, that Contarinia
spp. infestations reduced the level of
Megastigmus, but that the reverse did not occur.
The authors sug- gested that in North America
Megastigmus may be eliminated from the portions
of cones occupied by competing dipteran and
lepidopteran species, and that the increase in
levels of Megastigmus occasioned after some
insecticide treatments may be the reason that such
trials failed to result in increases in num- bers of
sound seeds.
Biological control
There have been investigations of insects that at-
tack seed and cone insects (Koerber 1960; Hedlin
1960, 1961a; Bringuel 1968; Miller 1983). Hedlin
(1960) reported that that Torymus sp. parasitizes
the larvae of Contarinia oregonensis and Glypta
eve- triae that of Barbara colfaxiana. He also
reported that many B. colfaxiana larvae perish
while attempting to be established in the cone,
with only about half of the original number of
larvae surviving lived beyond the first instar.
Other parasites of B. colfaxi- ana include
Tetrastigmus strobilus and Platymesopus sp.; these
chalcids were observed to kill 48% of B.
colfaxiana larvae. In one collection, 78% were
temperate-zone seed crops char- acteristically
reported a mortality rate for B. colfaxiana of support large parasite populations” (p. 481).
35%, virtually all caused by parasitoids and Several papers by Schowalter (1986, 1988, 1995)
predators (p. 1074). However, Roques (1991) discussed the interaction of forest manage- ment
noted that there are relatively few parasites techniques and the causes on possible con-
and predators of cone and seed insects (p.
303): “Parasite colonization thus apparently
requires adaption to both phytophagous host
and cone host” (p. 307).
One of the basic tenets of integrated pest
man- agement is that natural predators or
parasites of insects predacious upon crop
plants may be ef- fective controls of the pest
populations. There has been little evidence
that this concept has been em- ployed
successfully against seed and cone pests of
conifers (Yates 1989), and thus there has been
little discussion of the biological control of
cone and seed insects in the literature
reviewed. Exceptions insofar as Douglas-fir
seed pests are concerned in- clude the
following two organisms, which have been
shown to have potential for controlling
Barbara colfaxiana: Pyemotes, n. sp. (Acari:
Pyemotidae) and Trichogramma minutum
Riley. Moser et al. (1987) reported on new
species of Pyemotes that was para- sitic upon
Barbara colfaxiana as a possible natural
control of this seed pest. Hulme and Miller
(1988) conducted initial trials of
Trichogramma minutum in which it was
demonstrated that this parasitoid could
successfully attack eggs of Barbara colfaxiana
under field conditions; however, factors
possibly unique to this trial prevented the
intensity of parasitism nec- essary to control
Barbara populations. We found no additional
references to these two organisms relative to
cone and seed insects after these initial
reports, however. Keen (1958) noted that
parasites of the insect complex predatory on
Douglas-fir cones have been reared under
laboratory conditions. However, the cryptic
nature of the larvae of seed pests, as well as
the fact that much seed damage occurs prior to
parasitism, greatly limits the effectiveness of
such insects as control agents.
Other cases of mortality include micromammals,
the joint action of birds, rodents, and
entomopatho- genic fungi. Janzen (1971)
reported that “the insects preying on large
222 Douglas-fir: The Genus Pseudotsuga
trol of “pest” insect populations. Schowalter patholo- gists have always been aware of certain cone and
seed diseases, particularly cone rusts, it is only with the
(1986) noted that “silvicultural treatments can recent
promote or prevent insect population growth
depending on the ecological strategies represented
in the forest arthropod community. Forest
management or pest control strategies that fail to
address underlying causes of insect population
outbreaks or that fail to anticipate responses of
non-target insect species will be ineffective in
protecting forest resources in the long term” (p.
64). According to Schowalter (1995), “reduced
[insect] predator diversity increases the
probability that herbivores with rapid response to
environmental change will escape population
regu- lation by surviving predators” (p. 124).
Although he recommended natural forests as the
best overall approach, the foregoing references do
not detail the practical employment of such an
approach.

Diseases
Nelson et al. (1986) observed that “seed and cone
insects have long been known to reduce Douglas-
fir seed production. Loss in productivity from
fungi or other pathogens, however, has not been
measured, nor is it generally accepted that
significant losses from diseases occur in
developing cones” (p. 1). Shea (1960) noted that
“little attention has been given to fungi on forest
tree seed prior to sowing. Few seedborne diseases
of trees are known” (p. 2). Cooley (1983)
observed that “cone and seeds of conifers can
become diseased on the tree, during storage, or
when processed. Molds, a large diverse group of
fungi, and bacteria cause the majority of seed and
cone damage in the Pacific Northwest” ( p. 1).
According to Bloomberg (1966), Salsbury (1955)
“found that a high mold content did not necessar-
ily cause a reduction in the viability of Douglas-
fir seed (Pseudotsuga menziesii (Mirb.) Franco) and
it can be inferred that fungi exist endophytically in
seed without affecting the ability to germinate” (p.
413). Theisen and Goheen (1980) reviewed the
literature and observed that “not much
information is avail- able on the amount of
damage caused by cone and seed disease of
northwest conifers” (p. 1). Finally, according to
Sutherland et al. (1987, p. vi),
Although North American foresters and forest
advent of seed orchards, tree improvement pregermi- nation period. The early spring of
programs, and intensively managed forest
nurseries that the impor- tance of already-known
1958 was cool and wet, favorable to growth and
and newly discovered diseases has been development of seed mould. As environmental
recognized. A major technological change that conditions vary from year to year, so
has sparked interest in cone and seed disease has
been the change from collecting cones from wild undoubtedly does the relative importance of the
stands to the production of cone crops in seed biotic agents responsible for seed
orchards. There, the high value of the crop has
increased the importance of diseases. Sutherland
et al. (1987, p. vi),

Fungi
Perhaps the earliest report describing the
possible role of fungi with Douglas-fir seed
is that of Isaac (1935). This worker placed a
large number of seed in a rodent-proof
enclosure in the dense shade of a virgin
forest. Seed germination was recorded for
both the year of placement and the
following grow- ing season. The seeds that
failed to germinate were recovered and
examined. Many were found to be decayed,
but the fungus species involved were not
identified. In contrast, later studies designed
to eval- uate the effects of field stratification
and seeding date upon seed germination
utilized several thousand seeds placed in
rodent-proof enclosures (Lavender 1958a).
Subsequent tallies of field germination and of
laboratory evaluation of non-germinated
seed after the growing season demonstrated
little evidence of the destruction of viable
seed by fungi. However, again in contrast,
Lawrence and Rediske (1962), utilizing
seed treated with scandium 46, found that
fungi destroyed about 20% of the seed before
germi- nation, another 9% during the
germination period, and, finally, that
damping-off organisms killed yet another
9% of the seed placed after germination.
The difference between these results and
those of Lavender (1958, a-e) may have been
due to the much less vigorous field
germination of the seed employed in the
former study. Such response could reflect
ei- ther lack of stratification or weak seed.
Laboratory data has demonstrated that the
incidence of moldy seed in laboratory dishes
is inversely correlated with the vigor of the
seeds. Alternatively, as Lawrence and
Rediske (1962) suggested, “fungi accounted
for the greater part of seed loss during the
Chapter 9. Cone and Seed Insects and Diseases 223
maintained by these workers for 16 weeks at 20°C
loss” (p. 217). Laboratory evaluations of the fungi and 60% MC were covered with white mycelial mats
associated with Douglas-fir seeds (Salisbury 1953,
Holmes and Buszewicz 1955, Shea 1960,
Bloomberg 1969) have shown that Penicillium
spp., Aspergillus spp., Aureobasidium spp.
(Pullularia), Gliocladium spp, Spicaria spp.,
Trichoderma spp., and Mucor spp. are among the
more common of the over two score fungi species
isolated from Douglas-fir cones and seeds. In his
review, Shea (1960) noted that the first two
genera frequently are much more common on
seeds and cones in storage. Additionally, fungal
infection of the cones, especially those cones
gathered before the seeds were fully mature,
resulted in reduced viability of the seeds. In
extreme cases, when cones were stored under
particularly unfavorable condi- tions, fruiting
bodies of Aureobasidium pullulans (for- merly
Pullularia pullulans) were found on decayed
embryonic tissue. Shea suggested that part of the
loss in viability reported for seed stored above
freezing may be due to fungal attack. He further
noted that most seed was stored below freezing
and was free from active fungal attack during
storage. However, fungi have been shown to
retain viability “for at least 12 months when seeds
are stored at −12°C and ap- proximately 8%
moisture content. During this time no reduction in
germination was noted. However, seed carried
ample evidence of fungi which needed only
favorable conditions for continued mold growth
development on and in the seed” (p. 6).
Given the foregoing, much of the early work
investigating effects of molds on Douglas-fir seed
was concerned with cones stored prior to process-
ing. Lavender (1958) stored lots of cones
collected from 40 trees in the Willamette Valley.
Germination tests conducted with seed stored for
up to 4 months showed no effects of storage time
upon seed vigor. Unfortunately, no record was
made of cone moisture; however, the cones were
placed in small groups and probably dried during
the storage period in a manner similar to that
noted during other tests of cone storage. No
observations of mold growth on cones were made.
A more controlled experiment that demonstrated
the effects of fungi upon seeds stored under
conditions favorable to fungus growth was
reported by Rediske and Shea (1965). Cones
a function of the length of exposure to cool, damp
of Schizophyllum commune Fr. and yielded soil; the disease can spread from infected to
badly de- cayed seeds. However, cones held at healthy seed during stratification; it mummifies
0°C and 60% MC for the same period showed rather than rots the seed contents; and
only slight fungal attack. The authors also
reported that cones stored 40% MC or less
had little loss due to fungi after 16 weeks, so
apparently, unless the environment is
favorable to fungal growth, these organisms
do not infest Douglas-fir seed.
Bloomberg’s (1969) report was similar to
the above in that 125 days of storage in cones,
under conditions generally not very favorable
to fungal growth, resulted in little loss of seed
vitality, even though fungi that could attack
the seeds were pres- ent. But, when the seeds
were placed in germination dishes in an
environment apparently favorable to fungi,
significant numbers of seeds were diseased.
Bloomberg suggested that “the low
percentage of healthy non-germinable seeds
strongly indicates that the responsible fungi
and bacteria were facultative pathogens of
low viability seeds” (p. 180). According to
Bloomberg, the following genera were
common on Douglas-fir cones: Gliocladium,
Spicaria, Penicillium, and Trichoderma;
Rhizopus and Aspergillus were less frequent. A
basidiomycete, Corticium pini-canadensis was
often observed. The practical implication of
this research finding is that if mature seed lots
can be identified, they can be left for
extraction later than immature lots. The latter
may fall into two categories: those that
continually decrease in germinability, and
those that decrease and then regain some of
their original germinability. Other research
concerned with Douglas-fir seed disease has
studied Caloscypha fulgens and Fusarium spp.
Sutherland (1979) reported Caloscypha fulgens on
3% of Douglas-fir seed lots examined, but,
since it is a soil-borne pathogen, it was found
only in seeds from cones collected on the
ground, particularly in squirrel cashes.
Sutherland et al. (1987) described C. fulgens
as “an operculate discomycote with bright or-
ange (exterior often stained blue-green) cups
shaped fruiting bodies (1–5 cm in diameter). It
grows under conifers and fruits in the spring,
especially soon after snow melt” (p. 28). They
found that the degree of infection in seeds was
224 Douglas-fir: The Genus Pseudotsuga
it can survive for several years on the dead seed reported that bacteria (Pseudomonas chlororaphis)
(p. 31). Sutherland and van Eerden (1980) found could reduce the incidence of Fusarium spores found
that the problem with C. fulgens may “intensify on stratified
further if moist, stratified seeds are cold stored
prior to sowing. Additional spreading and killing
can occur following seed sowing, particularly
during prolonged periods of cool, wet weather”
(p. 14).
There has been considerable interest in the role
of seeds as carriers of a complex of fungi that are
among the most virulent of plant diseases,
Fusarium spp. However, of the early reviews of
Douglas- fir seed disease examined (Harvey and
Carpenter 1945; Shea 1960; Rediske and Shea
1965; Bloomberg
1966, 1969, 1970, 1973; Lock et al. 1975; Theisen
and
Goheen 1980; Cooley 1982) only Bloomberg
(1966) mentioned Fusarium. James (1986) and
Graham and Linderman (1983) reported low
levels of Fusarium on Douglas-fir seed, although
the latter suggested that the significant mortality
of Douglas-fir seedlings caused by Fusarium may
have come from seedborne propagules.
Sutherland et al. (1987, pp. 44–49, 52)
discussed pilch canker, Fusarium moniliforme var.
subglutinans, a major disease of southern pine, and
Fusarium oxy- sporum, a lesser problem of
Douglas-fir cones and seeds. Sutherland and
vanEerden (1980) noted that Fusarium
oxysporum causes root disease and top blight of
Douglas-fir seedlings but made no mention of it
on seeds. Nelson et al. (1986) reported a single
seedborne incidence of Fusarium on Douglas-fir,
but noted the following:
We consider recovery of only a single isolate notable,
however, because we sampled a broad range of
families on a broad range of sites over an entire
growing season. Finding only one isolate does not
mean that spores of the fungus could not have been
on tissue surfaces. If Fusarium spp. gain access to
nursery beds through infested seed, our results would
suggest that invasion of seed tissues occurs after cone
harvest and not while cones are developing on the
tree. (Nelson et al. 1986, p. 3)

We believe that seed and cone pathogens do not cause


significant losses in Pacific Northwest seed orchards,
but additional study of the pathology of cones in early
stages of development is needed. (Nelson et al. 1986,
p. 5)

In an interesting variation on Fusarium


research Hoefnagels and Linderman (1999)
Douglas-fir seed. James (1986) noted that the trees whose dormancy is interrupted, intact seeds
“extent of Fusarium contamination on seed are more re- sistant than damaged seed. This calls
varies greatly among conifer species and seed for extreme care in all phases of seed handling.
lots” (p. 268). Lori and Salerno (2002)
summarized literature concern- ing Fusarium
spp. and coniferous seed (including Douglas-
fir):
Seed-borne Fusarium may cause losses during
seed de- velopment, storage or germination and
damage may then result from loss of seed
viability or from seedling infec- tion following
germination. Fusarium may be found on and in
the seedcoat and in the gametophyte and embryo.
They may enter during seed and cone
development or through cracks in the seedcoat,
especially after the seed has been extracted from
the cone. External fungi on the other hand could
develop on the seed at any time after the seedcoat
is formed. (Lori and Salerno 2002, p. 560)

Bloomberg (1966, p. 419) found endophytic


fungi in the seed, but not in the embryo or
megagametophyte. Axelrood et al. (1995)
noted that “less than 2% of 67 Douglas-fir
seed lots from coastal British Columbia had
Fusarium and the incidence of individual
seeds varied from 0.3% to 95.4%. Running
water during imbibition reduced fungal
incidence over that found after standing
water” (p. 35). Allen (1947c, p. 51) noted that
disease-free germination of Douglas-fir seed
may be achieved in the laboratory through the
use of 30% hydroxy mercuric chlorophenol.
From the foregoing discussions, we may
conclude that much of the fungal growth
associated with seed is non-pathogenic—with
exceptions of but not lim- ited to Caloscypha,
Fusarium, Gliocladium, Trichoderma
Trichothecium, Cephalosporium, and
Aureobasidium; fungi generally are not a major
cause of seed mortal- ity in seed orchards; and
mature seed is more resis- tant to mold than is
immature seed (this is similar to the response
of spruce seedlings, wherein those that were
exposed to natural photoperiods prior to fall
lifting molded in storage, whereas those given
short photoperiods did not), and is further
evidence of the importance of annual
physiological rhythms to seed. And, if as
appears likely, global warming is affecting the
dormancy of Douglas-fir saplings, the current
virulence of Swiss needle cast (Phaeocryptopus
gaeumannii) may reflect reduced resistance of
trees occasioned by insufficient chilling, or in
10. Roots
Denis P. Lavender

A
lthough tree roots have been studied for over workers have suggested
a century, progress in understanding their
physiology has been slow, partly because until
about 50 years ago, studies were confined to the
anatomy and morphology of roots; partly because
roots grow in an opaque medium (soil), making it
difficult to conduct the necessary observations to
permit relating root growth to endogenous factors
and to components of the environment without
creating artifacts; and partly because root growth
is much more variable than shoot growth, and
therefore much larger populations are needed if
the data are to be significant statistically. In spite
of these quali- fications, there are numerous
reviews of tree growth and function (see, e.g.,
Hermann 1977, 2005; Sutton 1991; Fayle 1968,
1980; Reynolds 1975; Van Erden and Kinghorn
1978; Coutts 1987; Lyr and Hoffmann 1967;
Sutton 1980, 1990, 1999; Loescher et al. 1990;
and Comerford et al. 1990). Accordingly,
sufficient study of tree roots has been
accomplished to permit the following
generalizations, many of which apply specifically
to Douglas-fir, but which are probably equally
true in general, if not in detail, for the roots
of all coniferous trees.

Characteristics
1. Root growth is dependent upon carbohydrates
exported from shoots and other as yet
unidenti- fied substances, also exported from
the shoots.
2. Douglas-fir root systems do not develop dor-
mancy, although there remains much to learn
about this subject for Douglas-fir and other
trees. Hermann (1977) noted, “equally
disputed and unresolved is the question of
whether or not roots of trees in the temperate
climates become truly dormant” (p. 10). Some
that tree roots may have dormancy growth in the mild oceanic climate of England
(Lathrop and Mecklenberg 1971; Coutts become dormant as a strategy to reduce
1987, p. 763). If being dormant requires respiration. Fielder and Owens (1989, p. 543)
chilling to initiate ac- tive growth, found that some root activity occurred in
however, then Douglas-fir roots do not coastal but not interior Douglas-fir all year,
have true dormancy (Lavender et al. and that maximum growth occurred in fall and
1970). According to Hermann (1977), early spring. Individual roots may become
“observations of different cycles of root dormant and are generally unaffected by the
activity between different species under dormancy of the shoot, except that, when
similar environmental conditions would shoots are actively elongating, they attract
seem to strengthen the notion that at least most of the carbohydrate resources (are a
some degree of endogenous control exists, stronger sink) of the plant and, as a
however” (p. 11). consequence, root growth is diminished.
Krueger and Trappe (1967, p. 193) 3. While root growth may occur during the entire
cited two references noting that Douglas- year, except when soils are frozen or below
fir roots may elongate throughout the year. about 2 MP moisture tension, there are
Reynolds (1975, pp. 172–173), however, normally two periods of strong root growth
equated root dormancy with each year. The first, when most of the annual
metacutization of root apices, and, on this root growth occurs, is just before and during
basis, noted that the roots of Douglas-fir the period of bud break. The second,
somewhat smaller peak, occurs in

225
226 Douglas-fir: The Genus Pseudotsuga
the fall after rains have remoistened the soil support require a platform of heavy suberized roots
(Lyr and Hoffmann 1967, pp. 192–206). (McMinn 1963). As Coutts (1987) noted,
4. Root growth is affected by the environment of Although successful soil exploitation for the
the roots much more strongly than it is by absorption of water and nutrients requires a finely
divided root system, in trees the physical laws
shoot environment. governing the strength of beams under bending stress
5. The principal environmental factor which de- limit the degree of sub- division at the stem root
termines whether a root or stem will develop juncture commensurate with the firm platform
required for effective anchorage. In beams circular in
is pressure (Fayle 1968). cross section, stiffness is proportional to the fourth
6. Roots will grow in a temperature range of power of the diameter; thus, for effective support the
tree requires few but thickened roots at the base.
from 0°C to about 35°C, but maximum These develop by secondary growth of a limited
growth nor- mally occurs at soil temperatures number of primary roots (i.e., roots of primary
between 18°C and 25°C. structure) present on the seedling. (Coutts 1987, p.
761)
7. Coniferous seedlings generally have a limited
root elongation rate, rarely more than a very Such a design is obviously not efficient for water
few centimeters per day. In contrast, a single or nutrient uptake. To compensate for this
rye plant is capable of growing more than a deficiency, trees have evolved a symbiotic
kilometer of roots in a day. relationship with fungi, whereby the fungi
8. Most conifers are capable of forming a (mycorrhizae), with their very fine hyphae and
symbiotic relationship with one or more external enzymes, serve the function of fine roots,
species of mycor- rhizal fungi. Because the while the roots of the tree offer the necessary
fungal portion of the symbiosis has many support (Read 1991). Hunt and Fogel (1983, p.
times the absorbing area of the root and can 644) noted that the mean diameter of fungal
access soil particles and minerals that are not hyphae at about 3.0 to 4.5 μm in a Douglas-fir
available to roots, mycorrhizae are essential to forest in western Oregon; accordingly, these
the uptake of nutrients and water by the higher structures are well suited to penetrate the finest
plant. soil pores. As Finlay (2008) noted, “by virtue of
9. Data describing the annual turnover of root their small diameter the hyphae are also able to
biomass are limited, but it appears that a penetrate soil microsites which are inaccessible to
similar or even greater biomass of plant plant roots” (p. 1117). By comparison, fine tree
material is lost every year by root mortality roots have been characterized at 1 mm and small
than occurs in litter- fall. We describe the root roots at 1 mm to 5 mm in diameter (Santantonio
turnover in Douglas-fir stands later in this and Hermann 1985, p. 113).
section. The following presents a comparison between the
architecture of the very fine roots of a grass
Function (winter rye) plant (Dittmer 1937) and the
Tree roots have four major functions: relatively coarse, fine roots of trees (Table 10.1).
These data are not strictly comparable, as the rye
1. Uptake of nutrients and water plant figures show
2. Synthesis of organic compounds
3. Storage and transport of carbohydrates
Table 10.1 Root growth comparison between rye and Douglas-fir.

4. Anchorage of the tree and support of the aerial Root Root growth
Species length per day Reference
portion of the tree Winter rye 620 km 5 km Dittmer (1937)
The first to the third functions are definitely Douglas-fir 15 cm* Lopushinky and Max
compat- ible with each other and can be (1990) Trees in general 3- 56 mm Hermann
conducted in rela- tively fine root systems. Such (1977)
roots demand relatively 2 – 0 Douglas-fir
little photosynthate for growth and maintenance and are capable of exploring small soil
pores. In contrast, the great weight and stresses Total 15 cm Blake and Linderman
that tree roots must root/growth/ (1992)
seedling/day

* Per seedling
Chapter 10. Roots 227
Root regeneration potential
growth for the entire plant, whereas the tree data
are for individual roots. However, a seedling with The interest in root generation as a possible predictor
more than 100 elongating roots is rare. All data of seedling vigor and survival spread from the South
are for seedlings less than 3 years old. In
addition to the above the following tables
provide additional parameters of rye root growth
system produced in 4 months by a single seedling.
A comparison of these data with similar data
from conifers, where daily root growth is
measured in centimeters, shows the great
superiority of grass root growth. This same
relationship is evidenced by data describing
density of roots in the soil, i.e., 30–50 cm of grass
root per cm3 of the upper 15 cm of soil, as
opposed to 2 cm in the upper 8 cm of soil under a
Monterey pine stand (Barber 1977). Dittmer
(1938) found that a 16 cm3 soil sample from under
Kentucky bluegrass would have approximately
2,000 roots, 1.2 million root hairs, with a
combined length of over 372 m and a surface area
of about 419 cm2 (Dittmer 1938, p. 482). Given
the foregoing, it is clear that the mycor- rhizae,
whose extra-matrical hyphae may have a surface
area that dwarfs that of roots, is the major uptake
organ of tree roots and that the roots func- tion
primarily as support organs, translocate mate- rials
absorbed, and serve as a site for carbohydrate
storage and compound synthesis. In a comparison
of the cultivated rye plants grown in competition
with a non-competing greenhouse rye plant previ-
ously surveyed, it was found that the field rye had
approximately 5 times the number of root hairs
per unit of root length as the non-competing
greenhouse plant. However, the indoor plant had
far more and longer roots, and consequently a
greater total num-
ber of root hairs.

Relevance of Root Systems


for Seedling Survival and
Growth
Ever since Wakely (1949), stimulated by erratic
sur- vival of planted seedlings, introduced the
concept of varying physiological grades in
seedlings, research- ers have looked for
measurable seedling parameters that might define
seedling vigor and, hence, predict the survival
potential of a given seedling lot.
1. Slightly different terms have been used to refer to these measures;
to the West, where researchers at the in an effort to “overcome current confusion in the literature,” Day
(1982, p. 83) reviewed the use of “root regenerating potential” (RRP)
University of California (Berkeley) and “root growth capacity” (abbreviated “RGP”); Burdett (1987, p.
conducted a number of trials (Stone and 768) abbreviated “root growth capacity”as “RGC”; and Ritchie (1985,
p. 93) used the abbreviation “RGP” for “root growth potential.”
Schubert 1958, 1959a,c) on root regeneration
potential (RRP)1 with Douglas-fir and
ponderosa pine, and to Washington (Ritchie
and Dunlap 1980) and British Columbia
(Burdett 1979). Workers at Oregon State
tallied the incidence of new roots, but they
believed that those were a symptom of
seedling physiology and that seedling survival
was correlated with seedling dormancy (as
previ- ously discussed). Root generation was
the choice of a majority of investigators
because root systems apparently do not have a
dormancy cycle that could affect results and
because it was intuitively believed that rapid
root growth after planting was essential to
seedling survival. Accordingly, with the
exception of container-grown lodgepole pine,
which were subject to “toppling” (Burdett
1979), the inferred emphasis was on the role
of roots as absorbing organs. Wakely (1949)
hypothesized that (a) initial survival and
height growth of planted southern pines
depended on an excess of water intake over
water loss; and (b) the excess of water intake,
in turn, often depended on the formation of
new root tissue promptly after planting
(Sutton 1990, p. 260). We note here that the
many papers detailing root regeneration
potential generally do not mention
mycorrhizae, although there are numerous
papers detailing mycorrhizae formation on
coniferous seedlings (as will be dis- cussed in
a later section).
Ritchie (1985) and Ritchie and Tanaka
(1990) sug- gested that root growth capacity
(RGP) reflected the stress resistance of
Douglas-fir seedlings:
I am suggesting, therefore, that when we measure
RGP we are obtaining an estimate of relative
cold and stress resistance in the seedling and it is
these properties—not the ability to grow roots
per se—that influence how the seedlings will
perform on the site. A test of this hypoth- esis
would be to measure RGP, cold hardiness and
stress resistance over the course of a winter and
following different durations of cold storage. If
the relationship held up in the storage trials it
would seem to be valid.
228 Douglas-fir: The Genus Pseudotsuga
. . . RGP is a robust, relatively inexpensive and very concerning root volume and growth and suggested
flexible method for assessing seedling physiological
quality. (Ritchie 1985, p. 102) that such factors are not necessarily related to sur-
vival and growth. They noted in the discussion of
We agree—but it is difficult to understand why
their excellent study, that “the initial size and
the root growth should be maximal during
early post-planting expansion of root systems is
winter months, when roots generally do not
not pre- dictive of long term seedling (white
grow, and minimal in the early fall and spring
spruce) perfor- mance” (p. 1679). And Folk and
when Lyr and Hoffmann (1967) have found
Grossnickle (1997, p.
root growth maximal just prior to bud break.
121) noted that the predictive power of Douglas-
In contrast, Burdett (1987), working in root
fir seedling tests may be improved by conducting
growth capacity (RGC) tests, made the
them under unfavorable environments.
following points (numbering added):
In essence, RRP is the lifting of seedlings at
1. Published evidence on the relationship between lab- vary- ing times during the fall and winter, planting
RGC and field-RGC is virtually nonexistent. (Burdett
1987, p. 769)
them in containers, maintaining the containers with
favorable light, temperature, and water for 2
2. Thus the evidence seems convincing that root weeks, and then examining the seedlings for the
growth capacity plays a major role in determining the
ability of newly planted trees to avoid moisture stress number of new roots greater than a centimeter in
[by assuring good contact of roots with soil]. (Burdett length and, possibly, for the total number of new
1987, p. 769)
roots. Large numbers are interpreted to predict
3. It is probable, therefore, that moisture stress is nor- high survival. During the 1960s, when the pattern
mally a factor limiting establishment of planted seed- commonly reported was low survival of
lings. (Burdett 1987, p. 770)
seedlings and poor root growth in the fall (Stone
4. Thus early survival after planting in cold [limiting and Schubert 1959, a-c), this was true given fall
to root growth] soil is not necessarily unrelated to root
growth. (Burdett 1987, p. 770) vs. winter planting. More recent planting practice
has been to avoid late spring and early fall
5. Thus RGC tests may provide evidence of many
planting, when both the survival and root growth
types of plant injury” (Burdett 1987, p. 771)
were low, so that root growth capacity is no
6. Thus RGC does not predict survival but survival longer well-correlated with survival—because
potential. In general, the higher a seedling’s RGC the
greater its chance of surviving. (Burdett 1987, p. 773) survival then becomes more a matter of seedling
environment and not physiology. In 2000,
7. The RGC of forest tree seedlings measured under
standardized conditions in the laboratory often
Gourley and Lavender (unpublished data) noted,
predicts relative field performance. Evidence to show however, that Douglas-fir seedlings that had
whether this relationship can be explained by a received long-night treatments in later summer
correlation be- tween lab-RGC and field-RGC does
not exist. (Burdett 1987, p. 773) survived well after fall planting (see Chapter 8
for more detailed discussion of seedlings).
8. RGC tests provide a simple method of evaluating
the performance potential of forest planting stock,
Because of the great number of papers on this
which is probably cheaper than alternative methods of subject, we will limit our discussion to the follow-
similar predictive value. (Burdett 1987, p. 774) ing references, which either favor or criticize RRP
But, as was noted for previous reports, the above as a measure of seedling viability, and cite several
suf- fer by ignoring the probable effects of major reviews (i.e., Jenkinson et al. 1993, Sutton
mycorrhizae. Binder et al. (1990) presented data 1990, Richie and Dunlop 1980). According to
demonstrating that RGC can vary greatly with test Krasowski and Owens (2000), “it is obvious from
conditions, and summarized other publications the results that the initial size of the root systems
that agreed. They concluded that while RGC may did not predict post planting performance of
have great value in carefully controlled research white spruce seedlings and our study also
trials, it has definite weaknesses for project concluded that the initial growth of the roots in
plantings. Krasowski and Owens (2000, p. 1670) the first post-planting weeks did not predict
reviewed a number of reports seedling growth performance”(p. 1678).
(Seedlings in this trial did not encounter drought
stress.) In contrast, Nambiar (1981), working
with Pinus radiata seedlings, reported that
“the configu-
Chapter 10. Roots 229

ration and physiological state of the seedling roots which can be detected using a conductivity meter.
If the cell membrane is ruptured or the
developed in the nursery strongly influence the
transmembrane pro- tein pumps impaired, the cell
survival and speed of growth when outplanted” on contents leak at a greater rate. Electrolyte leakage
relatively dry sites in Australia (p. 118). These rate is therefore a measure of damage to the cell
membranes. (McKay 1992, p. 1372)
two reports emphasized the role of the
environment in root function. The methodology of REL involves measuring
Root regeneration potential has gained the quantity of electrolytes that leak from some
consider- able credence with regeneration workers sections of root prior to and after the sections are
because it may measure a trait considered vital to killed with heat; the higher values reflect increas-
seedling survival. However, no report known to us ing membrane damage. In general, a plot of these
describing root regeneration mentions the role of values over date of lifting had a strong correlation
mycorrhizal association in the phase of root with seedling survival for the same dates and with
growth most impor- tant to nutrient and water root regeneration data, although McKay and
uptake, i.e., mycorrhizae. Accordingly, root White (1997) noted that the strongest correlations
regeneration is possibly an indi- rect measure of were for seedlings out-planted on dry sites. They
seedling viability, but it is strongly influenced by found considerable variation between REL and
a number of factors (also including seedling Douglas-fir seedling growth and survival over a
dormancy), and, until we understand this range of sites: “A linear negative relationship
thoroughly, a true evaluation of root regeneration described the 3 sig- nificant relationships between
will not be clear. REL and Douglas-fir survival but there was no
In essence, root regeneration potential clear pattern in the form of the relationship
represents seedling vigor after a stress, i.e., lifting between REL and Douglas-fir growth” McKay
and any pos- sible storage, and not the and White (1997, p. 149).
endogenous root growth cycle. While many of Electrolyte leakage from fine roots is a robust
published trials show a cor- relation between the and easily measured parameter that has a rapid
root growth potential and field performance, as turn-around time and can be used to evaluate the
noted above, this is truly primarily for the high- viability of seedling root systems. REL measures
low extremes; and, as Ritchie (1994, 1985) noted, the ability of membranes within the root system
the data supported only a correlation and not a to contain ions. Damaged membranes tend to leak
cause-and-effect relationship. It is not clear ions so, if ion leakage is quantified, it can provide
whether seedlings do not produce roots and then an indicator of root viability. Ritchie and Landis
fail to survive because they lack new roots or that (2006) noted the following:
they fail to survive because they are not vigorous. REL has been used successfully to evaluate the effects
of cold damage, rough handling, desiccation, cold and
Generally, the correlation between dormancy and
warm storage, and other stresses on root viability and
survival is stronger than this correlation. seedling vigor.
REL is sometimes closely correlated with seedling
Root electrolyte leakage survival, but in other cases these correlations are
weak. This is because factors other than root damage
McKay and co-workers discussed the rationale for
can affect REL. Some of these factors are species,
the root electrolyte leakage (REL) technique and seedlot, seedling age, season, and bud dormancy
its usefulness in determining seedling quality in a intensity. When REL is calibrated for these effects it
can offer a simple, easy test of seedling root system
number of publications (McKay 1992, 1993, viability. (Ritchie and Landis 2006, p. 9)
1994, 1998; McKay and Howes 1996; McKay and
Mason 1991; McKay et al. 1993). McKay (1992) It should be recognized that like root regenera-
described the theory of this test as follows: tion, REL is strongly affected by seedling
dormancy, and, until we thoroughly understand
The movement of cell contents to and from cells is
the role of dormancy, we shall probably not know
con- trolled mainly by the structural proteins present
at points along the lipid bilayer of the cell the basis for either test. Ritchie and Landis (2006)
membrane [i.e., the transmembrane pumps]. When discussed this in detail, noting that REL varied
healthy tissue is put in water almost free of ions,
there is a slight leakage of cell contents, including
with species “and even seed source with species”
ions, into the surrounding water (p. 9). They cited the
230 Douglas-fir: The Genus Pseudotsuga
finding of Folk et al. (1999) that REL “must first Not withstanding significant expenditures during many
years of research and development effort, meaningful
be calibrated to bud dormancy status before it can
be effectively used to assess root damage in
Douglas- fir” (p. 9). This is additional evidence
that stress resistance in Douglas-fir is seated in the
dormancy state. They summarized the
effectiveness of REL as a predictor of outplanting
performance as follows:
The ultimate objective of any seedling quality test is
to predict how well nursery stock will survive and
grow after outplanting, and many studies have used
REL for this purpose. Unfortunately, results have
been mixed. With Sitka spruce and Japanese larch
seedlings, for example, REL was closely related to
both survival and height growth. In Sitka spruce and
Douglas-fir
seedlings, REL was correlated with survival on some
sites but not others (McKay and White 1997). REL
pre- dicted establishment of Japanese larch (Larix
leptolepis) seedlings to some extent, but Root Growth
Potential (RGP) was a better predictor (McKay and
Morgan 2001). Similar results were found with black
pine (Pinus nigra) (Chiatante and others 2002), while
Harper and O’Reilly (2000) reported that REL was a
poor predictor of survival potential in warm-stored
Douglas-fir seedlings. (Ritchie and Landis 2006, p. 8)

They noted further that REL is used in Europe but


rarely in North America.
We have discussed the two tests used most
com- monly to estimate seedling survival and
growth in Douglas-fir plantations. Other trials
have included seedling vigor tests, wherein both
control seedlings and stressed seedlings were
grown under favor- able conditions and speed
of bud break determined (McCreary and Duryea
1987), variable chlorophyll fluorescence,
tetrachlomide, and mitotic indices and stress
induced volatile emissions. Perks et al. (2001)
presented an interesting, useful discussion of
the possible role of chlorophyll fluorescence in
predict- ing Douglas-fir seedling vigor,
particularly after cold storage. All the tests
depended to some extent on determinations of
seedling dormancy. Perhaps the subject was best
summarized by Dunsworth (1997) and Van
Eerden (1994). As Dunsworth (1997) observed,
“for organizations that are currently us- ing
operational tests, adding new tests are likely to
result in diminishing returns since survival may
already be near optimum (90%) and large
growth gains from physiological testing are
unlikely” (p. 439). According to Van Eerden
(1994),
and readily measurable description of seedling temperature between 12°C and 18°C were
quality still elude nursery men and reforestation
personnel. As a result, seedling quality is largely
equivalent to similar changes in soil
described only in terms of crude characteristics, temperature in their effect upon seedling dry
principally morpho- logical parameters, unless weight” (p. 10). The
sound and practical seedling characteristics that
can be readily and inexpensively applied and
developed soon, of further research and
development efforts to characterize seedling
quality may become redundant and met with
expression of “So what?” (Van Eerden 1994, p.
67)

None of the tests consider the vigor and


nature of mycorrhizal component of the
root system. Obviously mycorrhizae are
extremely difficult to test meaningfully, but,
given the importance of the mycobiont
portion of the root system, it is perhaps not
surprising that investigations that ignore it
have not been more successful in estimating
seedling vigor.

Temperature (heat)
Nielsen and Humphries (1966) reviewed the
effects of soil temperature on root growth
and metabolism and noted “knowledge of
how root temperature affects plant growth is
woefully incomplete” (p. 5). Over the next
few decades, more studies began to ap- pear
that provided more information in this area.
In one of the few reports describing trials
wherein soil and air temperatures were each
independently con- trolled, Lavender and
Overton (1972) investigated the effects of
thermo periods and soil temperatures on the
growth and dormancy of Douglas-fir seed-
lings. They found that the best growth of
Douglas-fir seedlings grown from eight
provenances occurred at the highest of three
soil temperatures (10°C, 15°C, 20°C), and
that the soil temperatures had a greater
effect on root growth than did air temperature.
Some of their results are presented in Table
10.2, where each weight represents the
mean of 112 seedlings, 14 from each of 8
seed sources. The mean length of roots from
“dry” seed sources was 38 cm and the moist
was 33 cm. They reported that “a change of
1°C in soil temperature had a greater effect
upon both shoot and root dry weights of the
entire seed- ling population than did a
change of 1°C in day air temperature over
the entire range of air and soil temperatures
studied”; however, “changes in day air
Chapter 10. Roots 231

Table 10.2 Relation between ovendry weight of seedling shoots, roots, and entire plants to thermoperiods and soil temperatures.

Weight (mg) by soil temperature (°C)*


Shoot Root Seedling
Night temperature 20°C 15°C 10°C 20°C 15°C 10°C 20°C 15°C 10°C
Day temperature 30°C
24 475 389 261 255 232 186 730 621 447
18 499 374 279 265 197 153 765 571 433
12 471 351 304 285 270 195 756 621 499
6 518 355 251 260 187 204 774 542 455
Day temperature 24°C
24 602 327 250 301 182 161 903 509 412
18 390 368 278 251 187 189 644 554 467
12 430 347 227 233 215 166 663 562 394
6 446 311 262 238 207 168 685 51/8 431
Day temperature 18°C
24 497 332 216 218 192 196 714 524 412
18 465 300 224 227 210 173 692 510 397
12 397 300 215 234 189 211 631 489 426
6 406 263 221 181 161 229 588 425 450
Day temperature 12°C
24 308 248 195 229 213 166 537 461 361
18 260 212 190 209 187 181 470 399 371
12 201 175 127 181 174 143 383 349 270
6 194 177 138 203 183 171 397 360 310
* Each weight represents the mean of 112 seedlings, 14 from each of 8 seed
sources. After Lavender and Overton (1972, p. 10).

strong effect of soil temperature upon seedling dry Parke et al.


weight found in Lavender and Overton (1972)
agreed with the observations of Irgens-Møller
(personal communication), “who noted that
reduction of soil temperature from 20°C to 10°C
reduced both shoot and root growth of seedlings
grown from Douglas- fir seed collected in
Arizona and British Columbia when the plants
were grown under a constant 20°C thermoperiod
and 16-hour photoperiod” (p. 10). The results
sharply disagreed with the earlier data of
Steinbrenner and Rediske (1964), however. As
Lavender and Overton (1972) noted, Steinbrenner
and Rediske reported “a strong negative effect
upon both shoot and root weights with decreased
air temperatures, but a slight increase in shoot
weight and a small decrease in root weight with
lower soil temperature (Lavender and Overton
1972, p. 10).
Heninger and White (1974) found that soil
tem-
peratures between 15°C and 27°C were
optimum for root growth for a range of species.
(1983d, p. 660) reported maximum root
growth of Douglas-fir seedlings between
18.5°C to 24°C. Landis et al. (1993) found
that soil temperatures may affect container-
grown seedlings more than air temperatures.
Lopushinsky and Max (1990) tested the
effects of root temperatures between 0.5°C
and 30°C on the root growth of (Pseudotsuga
menziesii var. glauca (Beissn.) Franco). They
found that root growth began at 5°C, was at a
maximum at 20°C, declined at 25°C, and
ceased at 30°C. Plants of high-elevation
(1,372 m) seed source were lifted in mid-
March and grown in a greenhouse with a
22°C to 15°C and a 16-hour photoperiod for
5.5 weeks. (A very small amount of root
growth occurred below 5°C and at 30°C.)
Minore (1988, p. 217) found that Douglas-fir
roots were heavier in 16°C soil than in 8°C
soil when light intensity was about 420 Em-2
(20% full sun), but not when light intensity
was 1% of full sunlight. Krueger and Trappe
(1967, p. 197–98) noted that Douglas-fir
seedlings from Wenatchee, WA had
232 Douglas-fir: The Genus Pseudotsuga
some active roots at soil temperature of 1°C, but 31.0°C. According to McMichael and Burke
many more at 4°C to 5°C in the nursery. Peak root (1998), “in general, root growth tends to increase
growth occurred in July and March. Lopushinsky with in- creasing temperature until an optimum is
and Kaufman (1984) noted that there was no root reached above which root growth is reduced” (p.
growth of Douglas-fir seedlings during a 21-day 947). In their review, Kaspar and Bland (1992)
period at 0.2°C. Surprisingly, these temperatures stated that “root system expansion is a function of
are well below those (5°C to 8°C) for boreal two temperature- dependent processes, growth
species Pinus sylvestris and Picea abies. and development. Growth processes, like cell
Vapaavuori et al. (1992) and Hawkins et al. (1999, elongation, increase root length and diameter.
p. 61) found that fast-growing Douglas-fir Development controls duration of growth
seedlings had a lower shoot/root ratio, but not processes and initiation of new roots and
significantly so, than slow-growing seed- lings. reproductive organs” (p. 291). Much of the work
Reviews such as Lyr and Hoffmann (1967) and on tree roots has emphasized growth; however, as
Sutton (1991) emphasized that, in general, root they note, “further studies are needed to resolve
growth was more affected by soil conditions than uncertainties concerning the effects of
by temperatures and by moisture in the air. temperature on root diameter, root hairs, root
Hermann and Lavender (unpublished data) found turnover and root orientation” (p. 297). In
that this was certainly true for Douglas-fir. discussing the outplanting of seedlings, Sutton
(1994) observed, “soil conditions that are
Ambient soil temperature
particularly important are soil temperature and
In their review, Lyr and Hoffmann (1967) noted soil moisture, but any of a number of factors can
that tree root growth can occur under natural be dominant in any given situation.”
conditions between 2°C and 35°C. However, they
also stated that “it is difficult to give useful values Soil moisture
for mini- mum, optimum, and maximum Under natural conditions, it is difficult to separate
temperatures for root growth of trees. Most the effects of high moisture from low temperature
authors have not distin- guished between a be- cause wet soils are often cold. Lieffers and
physiological and an ecological optimum and Rothwell (1986) showed that under controlled
have neglected the influence of other factors on conditions, the roots of black spruce and eastern
these cardinal values. The method of measuring larch were more limited by anaerobic conditions
growth is very important in determining the with high water tables under high—rather than
temperature values. Therefore most data are not low—temperatures. Von der Gonna and Lavender
strictly comparable” (202). (1988) reported simi- lar results from field
Nightingale (1935) found that for apple and experiments with white spruce in British
peach trees, the optimum temperature for root Columbia; and Heineman and Lavender
growth was 19°C: “at 19° C and lower the newly (unpublished data) had similar results.
developed roots of both genera were typically Minore and Smith (1971) demonstrated
white, of rela- tively large diameter extremely significant differences among Northwest tree
succulent, lacking in mechanical strength and they species in their ability to grow roots over
characteristically exhibited few fine laterals and shallow water tables: lodgepole pine, red alder,
that at 24° C in both genera and the cortex turned Sitka spruce, and western redcedar all tolerated
brown and gradually sloughed off. The remainder shallow water tables, while Douglas-fir did not.
of the root, the central cylinder, was typically very In a lengthy review of root growth, Hermann
woody, of considerable strength and lacking in (1977) noted that “both lack and over-
succulence” (p. 633). abundance of water profoundly effects root
Douglas-fir roots grown when the entire seed- growth” (p. 13), citing references to trees,
ling was held at 4°C were also larger in diameter including Douglas-fir. According to Sutton
and succulent (Lavender and Waring 1972). Parke (1991), “the inter- related effects of soil
et al. (1983d, p. 658) found that Douglas-fir root moisture, soil drainage, soil aeration, soil
growth occurred at temperatures from 7.5°C to
fertility, soil temperature, and soil me-
chanical impedance, individually and
collectively are the prime determinant of root
system architecture”
Chapter 10. Roots 233
growth and shoot-to-root ratios were increased.
(p. 11). McCaughey and Weaver (1991) Thompson
demonstrated that Douglas-fir did not tolerate
submergence for more than 14 days at
temperatures between 13°C and 24°C. Lyr and
Hoffmann (1967) noted that studies on the
influence of low soil moisture on root growth
have been very infrequent. They also observed,
however, that “generally root growth decreased at
low moisture content,” and they quoted
Ladefoged (1939) to the effect that root growth
stops in most species when soil moisture is
reduced to 12% to 14% on an oven-dry basis (Lyr
and Hoffmann 1967, p. 207). Additional
considerations (and complications) are that soil
moisture is not uniform and that roots in moist
areas may absorb sufficient moisture to allow
roots in dry soil to continue growing. Lyr and
Hoffmann (1967) noted that root suberization
increased in dry soils, reducing the absorbing por-
tion of the root system.
Ritchie and Dunlap (1980) reviewed data for a
number of species, finding that root growth
generally fell with soil drying and that little
growth occurred at soil moisture tensions of –
1300kPa. Working with Douglas-fir, they found
that root growth decreased to soil moisture
tensions of –1500 kPa, and that stock listed in the
winter always made some root growth, even in
dry soils. They cited Stone (1967), speculat- ing
that “if new growth is just getting under way and
is potentially high when the seedling is planted,
the roots may continue to elongate when planted
in dry soil” (Ritchie and Dunlap 1980, p. 238).
Lyr and Hoffmann (1967, pp. 207-9) made the
following points relevant to this discussion:
• Root/shoot ratios were higher in dry soils.
• Lack of water inhibited root growth before
shoot growth.
• Root suberization increased in dry soils, hence
water absorption was less.
• Reduction of water, salt intake reduced
photosynthesis and hence root growth.

Light
Obviously, photosynthesis, which may vary with
level of light is essential to growth of all parts of
the plant, but, as Lyr and Hoffmann (1967, p.
218) found, shading primarily influenced root
rooting ability in Douglas-fir.
and Timmis (1978) reported that a number of
new roots of Douglas-fir seedlings increased Organic residues
with an increase in photoperiod from 8 to 12 As we mentioned earlier, ecosystems with ectotro-
to 16 hours. Van den Driessche (1970) noted phic mycorrhizae are characterized by a layer of
that seedlings of Douglas-fir and Sitka spruce
were independent on current photosynthesis
for root growth and that the root growth was
proportional to light intensity. Van den
Driessche (1987) utilized radioactive carbon
to demonstrate that root growth of seedlings
after planting in the spring was dependent on
current photosynthesis. Other work showed
that new root growth was proportional to
photosynthesis under a range of light
intensity, although a small amount of root
grew in complete darkness. Philipson (1988)
argued that “new root growth in Douglas-fir
plants is dependent on a living connection, the
phloem, with the shoot. This is consistent with
the view that in this species root growth
requires current photosynthate and possibly
other compounds translocated from the shoot”
(p. 106).
Zaerr and Lavender (1974) utilized girdling to
test the effects of materials exported from
the shoot on root growth of Douglas-fir
seedlings. They found that root growth was
absolutely dependent upon shoot exports,
and that food stored in the roots did not
substitute for current photosynthate, in
accord with Philipson’s later (1988) results.
In other trials, Zaerr and Lavender found that
photosynthesis levels in Douglas-fir
seedlings were not correlated with root
growth. Gilmore (1965) found that a
material from the shoot other than
carbohydrates stimulated root growth of
Loblolly pine seedlings. Minore (1988,
p. 219) found little root growth at 1% of
sunlight at either 8° or 16°C soil temperature.
Van den Driessche (1991) found that light
played no role in root initia- tion, other than
permitting photosynthesis, and that current
photosynthesis was not required. A small
amount of root growth was possible, probably
uti- lizing carbohydrates stored in the roots (p.
294). One-year-old seedlings could grow
limited roots in the absence of photosynthate,
but 2-0 seedlings could not. This may be
related to the rapid loss of adventitious
234 Douglas-fir: The Genus Pseudotsuga

Root growth Shoot growth


Root growth potential
Growth parameters
Dormancy cycle

Sep Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug

Dormancy True Quiescence Shoot Dormancy


deepening dormancy elongation induction

Figure 10.1 Root growth periodicity in Douglas-fir, with maximum and secondary cycles (after
Ritchie and Dunlap 1980, p. 224).

organic material (litter and humus) over the populations and should be composed of litter,
mineral soil, where most of the tree roots are humus and woody materials, all of which are
found. This is the major source of nitrogen and necessary to maintain soil biological material and
has the highest concentration of nutrients. It is productivity. Parke et al. (1983c) found that litter
where the majority of the microflora and enhanced the growth of Douglas-fir seedlings,
microfauna organisms that col- lectively maintain apparently through some biological effect other
soil production are found. Harvey et al. (1987) than that of mycorrhizae.
presented a detailed discussion of or- ganic matter Root growth periodicity
in the forest soils of the inland Pacific Northwest
Lyr and Hoffmann (1967) argued that data to sup-
and noted that 75% of the mycorrhizae and 96%
port a general pattern of root growth of trees does
of the roots in the Douglas-fir (Pseudotsuga
not exist, but noted that a period of maximum root
menziesii var. caesia) zone are found there. growth occurs in early summer, with a secondary
Harvey et al. (1987) recommended an organic peak in the fall. We have observed a similar
matter content of 31% to 45% in soils of the pattern in Douglas-fir (Figure 10.1), which was
Douglas-fir type. Such content is characterized by illustrated by Ritchie and Dunlap (1980, p. 224).
the highest mycorrhizal
11.Mycorrhizae
Denis P. Lavender

T
he famous line from English poet John Ecto
Donne’s 1624 “Meditation XVII” says that
“no man is an island, entire of itself.” In the
myc
introduction to orrhi
Les Symbiotes nearly two centuries later, the zae
French scientist Paul Jules Portier wrote, “All
Over
living beings, all animals from Amoeba to man, time,
all plants from Cryptograms to Dicotyledons are land
constituted by an association, the “emboîtement” plants
of two different beings” (Portier 1918, p. vii, as evolve
quoted in Lane 2005, d into
p. 14). This book is obviously concerned more
comple
primarily with Douglas-fir, but we digress here to
x
discuss the fungal components of the Douglas-fir
and
mycorrhizae. The term “mycorrhiza” is itself organiz
descriptive, as it is derived from the Greek and ed
means literally “fungus root.” Hacskaylo (1972) forms.
described mycorrhizae as the ultimate in reciprocal The
parasites (see also McDougall 1918, and Allen earliest
1991). This symbiosis is of a class that may be convin
cing
one of the major associations in botany, one that
evi-
is only slightly less significant than mitochondria
or chloroplasts. Indeed, the partnership is so close
that Wilhelm (1966) argued, “under agricultural
field conditions, crops do not, strictly speaking,
have roots; they have mycorrhizae” (p. 65). There
are three major types of mycorrhizae, which differ
primarily by the nature of the seed plants: (a) the
ericoid mycorrhizae, which occur primarily on the
dwarf shrubs of the heaths covering the fringe
areas of the world; (b) vesicular-arbuscular
mycorrhizae (VAM) on herba- ceous plants,
particularly in the lower latitudes; and
(c) ectomycorrhizae on trees of the Pinaceae
(pine),
Betulaceae (birch), Fagaceae (beech), and
Salicaceae (willow) families in forests of Europe,
Asia, and North America. Douglas-fir has
ectomycorrhizae, so we shall be concerned with
this form. 235
dence of VAM (vesicular-arboscular mycorrhizae)
is of a cycad dating from about 220 million years
before present, during the Triassic period,
although older records (350 million years BP) of
lycopod-like plants harboring vesicle-like
structures have been considered. Ectomycorrhizae
would have appeared later, likely prior to 70
million years BP, or perhaps in time to witness
the great cataclysm and the dinosaur decline,
some 65 million years ago. The principal nutrient-
absorbing organ of these symbioses are the
mycorrhizae, which are found primarily in the
litter layer that characterizes these forests. This
material, which has a C:N ratio lower than that
found under ericaceous plants is, nevertheless,
slow to decom- pose and hence form an acidic
layer enriched with organic material (Read 1991).
The litter layer varies from acidic more to mull,
which forms over base- rich soils. Research has
shown that some of the fungi have the ability to
mobilize nutrients from complex organic
molecules. Mycorrhizal roots have long been
known to take up phosphorus, but nitrogen is
frequently the limiting nutrient in these
ecosystems; fungi, such as the important Amanita,
Suillus, Boletus, and Thelephora have the ability
to produce an acid carboxypeptidase and hence
have the capacity to mobilize nitrogen from
protein (Read 1991). The pH range for these fungi
is 4.0–5.0, about the same as that which favors
litter layer accumulation. It is interesting and
important ecologically that the higher plant has
access to organic nitrogen only when it forms
mycorrhizae with an appropriate fungus.
Douglas-fir forms ectotrophic mycorrhizae with
about 2,000 non-species-specific mycobiants
(Trappe 1988). According to Pirozynski (1981),
“three events involving mycotrophic symbioses
appear to have been landmark events in the
progression of plant life on Earth: (1) the evolution
of endotrophic symbiosis,
(2) the evolution of ectotrophic symbiosis, and (3)
236 Douglas-fir: The Genus Pseudotsuga
the evolution of independence from mycotrophic 4.0 to about 7.0 and killed
symbiosis” (p. 1824). Pirozynski (1981) discussed
ectotrophic symbiosis in detail as follows:
The evolution of ectotrophic symbiosis predictably
had very different consequences because the
peculiarities of ectomycobionts contrast sharply with
those of endomy- cobionts. First, there are at least
5000 morphologically and physiologically distinct
species of ectomycorrhizal fungi (Eumycota, mostly
basidiomycetes and a few ascomycetes), suggesting
that the selective pressure has been on the
ectosymbiont which, in this case, is the mycobiont
forming a sheath on the outside of roots. Secondly,
ectotrophism has evolved only in some 2000 species
of plants, chiefly in Pinaceae, Fagaceae, Betulaceae,
Salicaceae, and within Dipterocarpaceae,
Caesalpiniaceae, and Myrtaceae. Thirdly, the
evolution of ectomycorrhizal symbiosis may be a
relatively recent event: the ectotrophs are selective in
more extreme en- vironments; the mycobionts
involved belong to more recently evolved “higher”
fungi; the fossil record of ec- totrophs extends back to
the second half of the Mesozoic; and the geographical
disjunctions of ectotrophic com- munities reflect
tectonic events of the same interval. The distribution
patterns imposed by these events are maintained by
relative immobility of ectotrophic systems, which
probably stems from the necessity for concomitant
dispersal of both components.
The evolution of ectotrophism did not result in an
major taxonomic discontinuity: in anatomy and mor-
phology the ectotrophs have not diverged significantly
from their endomycorrhizal progenitors and relatives.
But another glance at the ectomycorrhizal groups of
plants from the point of view of their habits and habi-
tats will reveal shared characteristics: they are woody,
arborescent plants which tend to form species-poor,
“pure stand” forests. In ectotrophic forests the
diversity at the root-soil interface appears to be
provided by the mycobionts. The dominant trees can
select as many as 2000 different species of
mycobionts according to to- pography, soil type, or
growth phase. Individual trees behave as
physiologically different organisms; they do not seem
to compete directly with each other and this may
allow them to grow in close proximity. Closely linked
with gregariousness is a frequent occurrence of
anemophily and of dry, unpalatable fruits.
To recapitulate, forest communities can be viewed
as mosaics of species with diversity deriving from
plants in one and from fungi in the other, being above
the ground in one and below in the other.
Furthermore, if differences in the composition of
forests have a mycotrophic basis, the pecularities of
each community may have evolved in consequence.
(Pirozynski 1981, p. 1826–1827)

Pirozynski noted that the mycorrhizal habit is


nec- essary for uptake of phosphorus and boron
in par- ticular. These two elements were found
to be low in hemlock. In one fertilizer trial,
hemlock saplings were fertilized with elephant-
grade urea, which raised the pH around the
mycorrhiza in surface litter layers from about
the mycorrhizae. As a result, the phosphorus layer to a multiple-layered compact covering. It can
have a variety of colors.
and boron content was lowered and tree Extraradical hyphae: loose, free-ranging hyphae
growth reduced (Gill 1981). Gill and extending outwards from the fungal mantle.
Lavender (1983a,b) argued that boron is Clamp connection: a small, semi-circular
connection between two cells of a hypha which
necessary for lignification and hence, the allows the transfer
mycorrhizal habit permitted the evolution of
woody stems. Malloch et al. (1980) observed
the following:
Those plants having endomycorrhizae usually
occur in forests of high species richness, whereas
those with ectomycorrhizae usually occur in
forests of low species richness. The roots of
ectomycorrhizal trees, however, support a large
species richness of fungal symbionts, probably
amounting to more than 5000 species world-
wide, whereas those of endomycorrhizal trees
have low fungal richness with only about 30
species of fungi known to be involved
worldwide. Ectomycorrhizal forests are generally
temperate or occur on infertile soils in the
tropics. The soils are characterized with an acid
humus layer. They have apparently expanded in
a series of ecologically important events through
the course of time from the Middle Cretaceous
onward at the expense of endomycorrhizal
forests. (Malloch et al. 1980, p. 2113)

According to Amaranthus (as quoted in


Wells 2000), “several miles of fungal
filaments can be present in a thimbleful of
temperate-forest soil. . . .
When they connect with the roots of plants,
they can increase the plant’s ability to take in
water and food by 10 to 1,000 times” (p. 22).
Certainly, the evolution of this relationship
has provided trees with absorb- ing capacities
comparable to that of grasses, and al- lowed
development of roots as primarily supporting
or anchoring organs. Interestingly, none of the
root regeneration papers discussed in the
section on roots has considered the role of the
presence or absence of mycorrhizae on
seedling survival and growth.

Anatomy
Massicotte (1994) noted that “there are two
main components to a mycorrhiza: the
fungus and the root,” and described the
anatomy as follows:
The Ectomycorrhizal Fungus
Hyphae or mycelium: fine thread-like cells
which comprise the body of fungal structures.
These can be a single hypha or they can be
packed together to form a mantle surrounding the
root or clustered together in a coherent structure
like a fruit-body, or hyphal strand.
Mantle: the layer of fungal hyphae covering
the root surface. It can range from a single, loose
Chapter 11. Mycorrhizae 237

of a nucleus to a new apical cell in growing the outer surface of the walls of the cells and
mycelium. It is a characteristic of Basidiomycete sometimes penetrate between two cells by separating
fungi which form ectomycorrhizae. the middle lamella. This network presents an
Septa: cross walls which divide fungal cells. organization similar to the one of the ectomycorrhizal
Hartig net: labyrinthine branching of fungal cells Hartig net . . . (Dexheimer and Pargney 1991, p. 312)
found between root epidermal and cortical cells in an
Ectomycorrhizas comprise a mantle of hyphae around
ectomycorrhiza. The multiple branches derive from
the root . . . and a network of intercellular hyphae, the
fungal cells in contact with the root cells.
Hartig net, . . . in which the symplastic continuity be-
The Root tween the cortical host cells is maintained. The Hartig
Apex: the growing tip of a root. net constitutes an important area of contact between
the fungus and the root cortical cells. ATPpase
Lateral root: a root branch which has derived later-
activities, demonstrating active transport, have been
ally from another root.
localized on the plasmalemma of both partners,
Root hair: a hair-like cell extension radiating out-
and most
wards from a epidermal cell on the root surface.
exchanges are considered to take place in this part of
Root cap: a collection of root cells covering the
root apex. the mycorrhiza. (Dexheimer and Pargney 1991, p. 317)
Apical meristem: the zone of dividing cells at the Dexheimer and Pargney (1991) noted that, although
root apex which give rise to new cells in a growing
root. Epidermis: the outer most layer of cells of the there are different kinds of mycorrhiza,
root and the ones indirect contact with the soil micro The interfaces between the symbiotic fungi and the
envi- ronment (or covered by fungal portion of cells of the host plant are always bordered on the one
mycorrhizae). Hypodermis: the layer of cells side by the fungal plasmalemma, and on the other by
underlying the epi- the plasmalemma of the host plant or the
dermis (fungal hyphae extend between cells). persisymbiont membrane derived from it. The
Exodermis: the layer immediately beneath the epi- cytoplasms of the two partners never come into direct
dermis, but only called exodermis when the cell walls contact and are sepa- rated by a mixed apoplast
contain a Casparian band. Some deposit suberin comprising a fungal wall and a part originating from
lamel- lae as well. the host plant (wall or isolation layer). However, the
Cortex: the cells between the epidermis and interfaces of the endomycor-
endo- dermis (fungal hypha are not present). rhizas and ectomycorrhizas are not exactly identical.
Endodermis: the cell layer surrounding the In ectomycorrhizas, the part derived from the host
vascular cylinder in the middle of the root. (Fungal plant is the cell wall (Dexheimer and Pargney 1991,
hyphae do not invade this tissue.) p. 318)
Vascular cylinder or stele: the zone internal to the
endodermis which contains root vascular tissue
(xylem and Phloem).
Mycorrhizae
Pericycle: the cells immediately interior to the Parke et al. (1983b) noted that ectomycorrhizae
endo- dermis which give rise to lateral roots and part
of the vascular cambium. are often assumed to increase water uptake of
Apoplast: the zone outside of the plasmalemma of conifer- ous hosts. They presented a detailed
cells within the root. study of the effects of mycorrhizae during drought
Symplast: the continuous zone inside the plasma-
lemma of living root cells of the entire root. periods and concluded that at least one fungus
Middle lamella: the zone of pectic substances con- (Rhizopogon vini- color) reduced transpiration of
necting root cells together. Douglas-fir seedlings and increased water uptake,
Plasmodesmata: the microscopic connections be-
tween the symplast of adjacent root cells. and that these seedlings recovered faster from
Intercellular space: spaces outside the root cells drought than did non-my- corrhizal plants and
usu- ally in the cortex at the junction of cells.
seedlings with different fungi. Castellano and
(Massicotte 1994, p. 12)
Trappe (1985, p. 616) succeeded in inoculating
Dexheimer and Pargney (1991) described the Douglas-fir seedlings with Rhizopogon vinicolor.
interface of vesicular arboscular mycorrhizae in After 2 years in the field, survival of those
two passages as folllows: seedlings was 11% higher than survival of non-
The fungus produces a network of intercellular inocu- lated plants, and the inoculated seedlings
hyphae within the root cortex. From the hyphae of
this network, branches penetrate the cell wall and were 245% larger. Bledsoe et al. (1982) found that
form ramified intra- cellular structures, the arbuscules incubating Douglas-fir seedlings with Hebeloma
Often the hyphae
crustuliniforme or Laccaria laccata did not
dilate to produce ampoules with thickened walls, the
vesicles. In most VA mycorrhizas, the hyphae of the increase survival on a dry site east of the
intercellular network are located in the spaces Cascades. The mycorrhizal fungi did not compete
between the cortical cells where they are in close
contact with
well, reduced seedling biomass, and, since they
were from western Cascades isolates, were
probably not well adapted to the planting site.
238 Douglas-fir: The Genus Pseudotsuga
Parke et al. (1983b) reported that Rhizopogon been unable to successfully inoculate seedlings with
great- ly increased the drought tolerance of fungi in S.W. Oregon” (pp. 89–90). Owston et al.
Douglas-fir seedlings, noting that the “net (1992) concluded,
photosynthetic rate of Rhizopogon inoculated
seedlings 24 hours after rewatering was seven
times that of non-mycorrhizal seedlings. The
transpiration rate of Rhizopogon inocu- lated
seedlings was low before dessication, declined
rapidly during the drought period, and, after rewa-
tering, quickly resumed a rate higher than for
other treatments” (p. 83). Molina et al. (1999)
noted in a review paper that “Rhizopogon is the
largest genus of hypogenus Basidiomycota with
worldwide distribu- tion among Pinaceae. They
occur in young and old forest stands alike and in
diverse habitats” (p. 129). They also noted that
this fungus produces prolific rhizomorphs that
play an important role in water uptake by
seedlings (p. 153). Dosskey et al. (1990)
demonstrated greater tolerance to drought when
Douglas-fir seedlings were infected with R.
vinicolor. Molina et al. (1997) found that
Rhizopogon can colo- nize both Douglas-fir and
manzanita (Arctostaphylos) and that this may
facilitate the survival of Douglas-fir under
droughty conditions.
Dosskey et al. (1991) reviewed a number of
pa-
pers indicating that mycorrhizae can increase
seed- lings’ tolerance to drought and presenting
data that demonstrated that Rhizopogon-infected
(but not Hebeloma crustuliniforme or Laccaria
laccata) mycor- rhizal Douglas-fir seedlings had
increased photo- synthetic rates under drought
stress. Rhizopogon had no effect on water uptake,
reduced the length of roots, and enhanced the
stomatal conductance, but reduced leaf-water
potentials (p. 332). Duddridge et al. (1980)
discussed the rhizomorph function in the uptake
and transport of water and demonstrated that
mycorrhizal pine seedlings survived for several
weeks under drought conditions that killed the
non- mycorrhizal control seedlings.
Molina (1981) reported that seedlings
inoculated with Pisolithus tinctorius survived
better on hot, dry sites in southwest Oregon and
that “early plantation mortality in Southwest
Oregon is not attributable to root disease but is
related to deficiency in mycor- rhizae formation
and function. Unfortunately, many trials have
Rhizopogon vinicolor is the best candidate for −2.0 Mpa corresponded to an increase in absciscic
inoculation of nursery seedlings. It is easily acid content of 240 ng g−1. The relationship between
inoculated, aggressively colonizes roots and abscisic acid and water potential was not definitive,
competes well against other mycor- rhizal fungi, though the general trend was an increase in the
and produces abundant quantities of hyphal hormone with intensifying stress until water
strands that extend long distances into the soil to potential was −5.0 Mpa, when concentration
help seedlings take in water and nutrients. sharply declined. (Johnson and Ferrell 1982, p. 431)
Furthermore, R. vinicolor persists and spreads on
root systems of seed- lings after outplanting so We have made a number of references to
that seedlings can continue to benefit for several mycor- rhizae and stress in the material covered
years” (Owston et al. 1992, p. 322).
thus far. The importance of the mycorrhizal habit
Coleman et al. (1990) found that infection in establishment and survival of forest trees is not
of Douglas-fir roots with any one of several in doubt. Which fungi stimulate particular tree
fungi did not affect root hydraulic species under specific abiotic and biotic
conductivity. Further, this result was not conditions is still an open question. Most biotic
affected by phosphate levels. There was no and abiotic parameters are not studied rigorously
relationship between seedling levels of either enough to allow coherent predictions across a
zeatin riboside and abscisic acid and root wide range of conditions.
conductiv- ity. Johnson and Ferrell (1982) Considerable research on effects of ectomycor-
studied abscisic acid and its metabolites rhizal inoculation on outplanting performance of
through two drought cycles with Douglas-fir forest seedlings continues. Unfortunately, most of
seedlings: the current work is similar to previous work using
Three year-old intact seedlings were water- species of Pinaceae as the hosts and Pisolithus
stressed, watered, and restressed over a period species as the inoculated mycorrhizal fungus.
of 30 days. . . . Additional attention needs to be focused on native
Needle content of absicsic acid and 2-trans-
absciscic acid and their saponifiable conjugates plants and host-specific mycorrhizal fungi.
were quantified with gas-liquid chromatography. Although the ap- plicability of the results are not
The typical water potential threshold in branch as wide ranging, the potential for outplanting
conductance, decreasing abruptly at
improvement may well be higher.
Chapter 11. Mycorrhizae 239

Summary ecological role of these symbionts is “in protecting


The reports cited demonstrate that some fungi in southwestern Oregon. They also noted that
mycorrhi- zal fungi may increase the survival of disturbed soils from clearcuts and soils from old
Douglas-fir seedlings under drought stress; growth stands had equivalent my- corrhizal inoculum.
however, the spe- cies and ecotype must be Parke et al. (1983a) found that there was less
chosen as carefully as the seedling source. mycorrhizal inoculum in clearcuts, whether burned or
not, than in undisturbed forest soil. However, there
Ecology was sufficient mycorrhizal in- oculum in burned soils;
Parke et al. (1983c) reviewed the literature thus, common plantation failures in southwestern
concerning the role of forest litter in the growth Oregon were probably not due to lack of mycorrhizae
forest trees and noted that its effects on soil on the seedlings. Wright and Tarrant (1958) found
chemistry and physical state increased site that, in the Wind River area of southern Washington,
productivity: “Litter amend- ment usually the occurrence of mycor- rhizal roots of Douglas-fir
enhanced growth of host seedlings but growth was inversely correlated with degree of burning;
enhancement could not be fully attributed to however, this was not true on the H.J. Andrews
addition of mycorrhizal inoculum or nutrients Experimental Forest in the Oregon Cascades.
provided by litter. These findings suggest that Mycorrhizae were not found on roots in decayed
other biological factors stimulated growth of wood or in rocky or compacted soil. There was no
conifer seed- lings and (or) activity of relation between soil pH and incidence of
mycorrhizal fungi” (p. 666). Parke et al. (1983d) mycorrhizae. As a group, mycorrhizal fungi are found
found that the soil temperature range of 18°C– over a wide range of soil environments, but individual
24°C was optimal for the growth of mycorrhizal fungi are strongly influenced by soil moisture,
temperature, pH, fertility, and organic matter. As ecosystems in the final stages of succession,
noted previously, ectomycorrhizal forests on a where they keep nutrient cycles closed and
worldwide basis are characterized by acid litter prevent loss of resources from the entire system”
layers that decompose slowly. (p. 312). Simmard et al. (1997) found that, while
Pankow et al. (1991) briefly reviewed the ecology mycorrhizae did not increase the growth of
of mycorrhizae and suggested that the principal Douglas-fir seedlings, “the influence of overstory
trees and pattern of ectomy- corrhizal formation
are important to P. menziesii seedling
performance in deeply shaded forest en-
vironments” (p. 327). Trappe (1988) argued that
the classification of mycorrhizal fungi ignored the
physiology and ecology of fungi and
recommended instead a classification of
mycorrhizae, “by mycorrhi- zal dependence of
hosts as it interacts with dispersal strategy of the
[fungi]” (p. 347).
Mexal and Reid (1973) found that both
flooding and drought were limiting factors in
ectomycorrhizal formation. They reported that
Cenococcum graniforme made maximum growth at
–15 bars moisture, where- as Thelephora
terrestris failed to grow and Suillus luteus grew
poorly: “While some species may grow or even
thrive at ψ approaching −40 bars, the spe- cies
studied here were severely limited by ψ below
−15 bars” (p. 1584). Cordell and Marx (1994)
noted that coarse, well-drained soils promoted
mycor- rhizal growth and that the pH range
favored by the host plant also favored the fungi;
high soil organic content also favored the fungi,
while excessively high nutrient content of
fertilized soils inhibited the development of
mycorrhizal fungi.
Parke (1985) observed that “mycorrhizal fungi
occur in a large diversity of plant communities
and their adaption to extremes of environment is
widely acknowledged” (p. 107). According to
Trappe (1977), “temperature profoundly influences
growth, me- tabolism and colonization of roots by
mycorrhi- zal fungi” (p. 211). Husted (1991) and
Husted and Lavender (1989) found that soil
temperatures of 15°C–17°C were more
advantageous to mycorrhizal fungi endogenous to
northern British Columbia than were temperatures
of 6°C or 26°C. Trappe (1977) noted that
Pisolithus tintorios grew well at 40°C and that
“temperature is an important parameter of
mycorrhizal growth” (p. 211). Dighton (1991)
citing Read (1984), noted the following:
Changes in vegetation, soil organic matter
accumula- tion and mycorrhizas [occur] on both an
altitudinal
240 Douglas-fir: The Genus Pseudotsuga
and latitudinal gradient. [Reid’s] scheme shows that Fusarium by extracts of L. laccata over a range of
as one ascends in altitude or progresses toward the
poles from the equator, in general one moves from conditions and found L. laccata was effective, thus
soils which are mainly inorganic (due to rapid antibiosis under
decomposi- tion and nutrient cycling) through soils of
increasing organic matter content (due to lower
decomposition) to very poor, shallow and mainly
inorganic soils (due to limited plant productivity in
climatically adverse conditions at extreme altitude or
close to the poles). .
. . Consequent changes in mycorrhizal flora are
from vesicular-arbuscular mycorrhizas through the
ecto- mycorrhizas to the ericaceous mycorrhizas.
Thus, it is expected that the ability of mycorrhizas
to degrade organic matter and become involved in
direct cycling of nutrients is of major significance in
the ectomycorrhyzal and ericaceous mycorrhizal
dominated communities. (Dighton 1991, p. 363)

Parke et al. (1983c) found that forest litter,


particu- larly from undisturbed forest stands,
stimulated seedling and mycorrhizae growth.
They suggested possible saprobism for
mycorrhizae.

Disease
Linderman and Hoefnagels (1993) noted that
“while mechanism to reduce root diseases (by
mycorrhyzal seedlings) is not clearly
understood, it appears to involve some
morphological or physiological chang- es that
occur when the mycorrhizal association is well
established.” The authors emphasized the im-
portance of mycorrhizal-caused changes in root
membranes in the formulation of bacteria
colonies adjacent to roots in controlling
Fusarium spp., and described the development
of studies designed to identify soil
microorganisms antagonistic to Fusarium and other
seedling root diseases.
In a series of papers, Sinclair and coworkers
in- vestigated several methods of reducing the
pathol- ogy of Fusarium oxysporum on Douglas-fir
seedlings. Stack and Sinclair (1975) investigated
the ability of two common mycorrhizal fungi in
Pacific Northwest nurseries, Laccaria laccata and
Inocybe lacera, to reduce Fusarium oxysporum
infection. They found that L. laccata, but not I.
lacera, offered some protection. In a further trial,
L. laccata provided protection before mycorrhizae
were formed. It was shown that when
L. laccata was placed between the Fusarium and
the seedling, infection was reduced, but the reason
was not clear. Sylvia (1983) and Sylvia and
Sinclair (1983a) investigated the inhibition of
laboratory conditions was demonstrated. since several appear to be inseparable (i.e., mantle
These workers examined the roots of barriers, host origin inhibitors, differences in
Douglas-fir seedlings (Sylvia and Sinclair chemical exudations, etc.). This broad spectrum of
1983b) and found that phenolics induced in defense mechanisms acting in concern assures
the cortical tissue by L. laccata, and not greater opportunities for biological control of
antibiosis, was the mechanism for the feeder root pathogens by my-
reduction of Fusarium pathogenicity.
Additionally, Sinclair et al. (1982) showed
that L. laccata stimulated greater seedling and
shoot/root growth. In a later publica- tion,
Strobel and Sinclair (1991a) showed that L.
bicolor stimulated flavonolic infusions in the
cortical cell walls that restricted fusarium to
intercellular spaces. In a later report, Strobel
and Sinclair (1991b) noted that, in previous
trials, “a high degree of root protection
[prevention of lesion formation] was ob-
tained against a relatively non-aggressive
isolate of
F. oxysporum at a temperature below optimum for
disease development” (p. 420). They
concluded: “The timing and intensity of wall
infusion with phenolics may be crucial
determinants of resistance of Douglas-fir to F.
oxysporum Unfortunately,
the highly specific nature of the requirement
for this induced resistance and variable
expression of induced resistance mechanisms
appear to preclude its practical exploitation”
(p. 424).
Marx (1973) reviewed a large number of
papers, some of which presented evidence that
mycorrhizal fungi may protect roots against
pathogens. The ef- fects varied with
mycorrhizal fungi and pathogens and may
have been the result of the physical barrier
posed by the fungus, the production of
biocides by either the fungus or the host plant,
or the develop- ment of a bacterial population
in the mycorrhizal rhizosphere. Finally, he
cited Zak (1964), who sug- gested that the
mycorrhizal fungus may utilize the
carbohydrates that would otherwise attract
and feed pathogens.
Sinclair et al. (1982) demonstrated that
Laccaria lac- cata protected Douglas-fir
seedlings against Fusarium oxyparum even in
the absence of mycorrhiza. Marx (1973)
concluded: “In all probability most of the
proposed mechanisms of root protection by
mycor- rhizae are functional at any given time
Chapter 11. Mycorrhizae 241
mechanisms could be involved, but changes in the
corrhizae” (p. 877). Unestam and Damm (1994, p. microbial populations in the miccorhyizosphere seems to
be the best explanation, yet least studied. . . .
175) largely agreed with Marx’s hypothesis,
noting that long roots on mycorrhizal seedlings,
which are non-mycorrhizal, suffered less disease
than those of non-mycorrhizal seedlings,
suggesting that the acid environment strongly
inhibited nursery pathogens. They also stated,
however, that “the work on mycor- rhizae as a
control agent has been conducted entirely on a
basic level with tree seedlings, no apparent
applications have yet been developed for use in
the nursery and doing outplanting in the field. The
reason for this is the lack of knowledge on fungal
behavior in different soils, since the methods
work elegantly in some environments but root
protection is often difficult to predict and not
ubiquitous” (p. 173).

Mycorrhizosphere
Linderman (2000) noted that the “rhizosphere”
phe- nomenon was first discussed in 1904. In a
number of papers, Linderman (e.g., 1985, 1988,
2008) extended the concept to include
mycorrhizae, terming the phenomenon “the
mycorrhizosphere.” He presented data supporting
the concept of a community of soil
microorganisms “wherein roots attract
mycorrhizal fungi and the latter attract bacterial
associates.” Chanway (1997), Chanway and Holl
(1992, 1994) and Chanway et al. (1991a)
discussed soil bacte- ria in the Douglas-fir
rhizosphere that stimulated seedling growth. One
major effect of this group of soil microoranisms is
to suppress root pathogens. Linderman (2000)
described this concept in detail, crediting it with
the general lack of root disease in natural soils.
Linderman and co-workers published a series of
reports that implicated the soil microbial
populations formed as a result of mycorrhizal
infec- tion with the concurrent change in exudates
from the mycorrhizae with a definite role in
enhancing the host plant’s health and vigor
(Linderman and Hoefnagels 1993). Linderman
(2000) presented sev- eral hypotheses as a cause
of this effect:
(a) enhanced nutrition, (b) competition for nutrients
and infection sites, (c) morphological changes, (d)
changes in the chemical constituents implant tissues,
(e) alle- viations of abiotic stress, and (f) microbial
changes in the mycorrhizosphere. Depending on the
disease and environmental situation, any or all
several weeks after control seedlings died. They
The mycorrhizal association with roots of
land plants has existed for hundreds of millions
noted that rhizomorphs are necessary to pro-
of years and logically includes associations with vide increased seedling survival. Parke et al.
other functional groups of soil microbes (1983a) reviewed a number of studies
Currently, we observe that plants have little
or no disease, or at least no strong impact of
demonstrating that mycorrhizal coniferous
disease on their growth and survival, in natural seedlings tolerated drought
ecosystems where microbial balance (including
mycorrhizas) has not been disturbed. (Linderman
2000, p. 345-346)

Of course, Linderman was concerned with


soil diseases. Unfortunately, although there
have been numerous attempts to draw
conclusions about my- corrhiza-disease
interaction, the lack of data has limited such
approaches. Linderman (2000, p. 359)
concluded that “the primary mechanism of
mycor- rhiza-disease interactions is through
the induction of changes in the microbial
community” (caused by altered root and
hyphal exudations).

Nutrient and water uptake


In addition to exploring great volumes of
soils, the small diameter of fungal hyphae
allow wider pen- etration of soil. Mycorrhizae
have external enzymes that enable dissolving
phosphorus from rock and produce chelators
that prevent the phosphorus from binding with
other elements. Accordingly, they are much
more efficient at taking up phosphorus than
are higher plants. Nutrients, particularly
phospho- rus and boron (Gill 1981), are
transferred to the host plant, while the fungus
derives carbohydrates from its host. Trappe
and Strand (1969) reported the stunting of
non-mycorrhizal Douglas-fir seed- lings in an
Oregon nursery. They found that severe
phosphorus deficiency in Douglas-fir
seedlings was associated with a lack of
mycorrhizae. Rygiewicz and Bledsoe (1984)
reported that mycorrhizae ap- peared “to
enhance K uptake and storage in roots by
increasing the vacular pool sizes, increasing
influx rates and decreasing efflux rates” (p.
123). Perry et al. (1987, p. 929) stated that
“mycorrhizae improve seedling survival and
growth by enhancing uptake of nutrients
(particularly phosophorus).”
Duddridge et al. (1980) demonstrated
that water flow in fungi rhizomorphs was
sufficient to enable Pinus sylvestris
seedlings in a dry soil to remain vi- able for
242 Douglas-fir: The Genus Pseudotsuga
better than control seedlings did; furthermore, synthale from one host species (Betula papyrifer)
they recovered more rapidly after drought stress. to Douglas-fir under field conditions. Noting that
Parke et al. (1983b) compared mycorrhizal and non- the results were similar to those of a number of
mycorrhi- zal Douglas-fir seedlings during and studies in laboratory conditions, they suggested
after drought stress. The performance of seedlings the follow- ing: “Consequently, the theory that
inoculated with Rhizopogon vinicolor was plant community dynamics operate mainly within
superior to that of control seedlings or those the constraints of resource supply should be
inocculated with other fungi. A second fungus reformulated to consider mutualism between
that has been reported to tolerate low soil plants and their mycorrhizal fungi, as well as
moisture is Cenococcum geophilum. In gen- eral, microbial resource sharing” (p. 520). Hacskaylo
however, mycorrhizae develop poorly under (1973a) reviewed the carbohydrate metabolism of
moisture stress. Further, most mycorrhizal hyphae a range of sugars by a range of fungi and warned
are in the upper soil layers and do not contact the against generalizations, since the data were
moisture that may be in the deeper portion of soil erratic. As he wrote, “In nature, ectomycorrhi- zal
horizons. According to Trappe (1977), “Of all the fungi depend primarily upon the roots of their
ectomycorrhizal fungi, Conococcum geophilum is hosts for carbohydrates, usually sucrose, glucose,
best recognized as drought resistant” (p. 212). and fructose. Certain species of fungi may,
Trappe also noted that Hymenogaster alnicola and however, possess enzymes to hydrolyze cellulose
lacternius obscorutus tolerate very wet soils. and other carbohydrates, but this characteristic
does not appear to be widespread” (p. 227). Some
Physiology ectomycorrhizae fungi have the capacity to covert
Scagel and Linderman (1998) noted that mycor- the foregoing sugars to mannose, trehalose, and
rhizae formed in Douglas-fir by the fungi, Laccaria glycogen, which are not
laccota and Rhizopogon vinicolor increased the utilized by the host phytobiont.
root concentration of Indole-3-acetic acid (IAA).
The mycorrhizal seedling height, diameter, and Soil
shoot/ root ratio were all increased and correlated In greenhouse trials, Borchers and Perry (1989)
with the IAA concentration in the roots: “In several dem- onstrated that soil collected under
cases, the extent of colonization was correlated hardwoods fa- vored mycorrhizal and seedling
with in vitro IAA or ethylene production capacity growth. Amaranthus and Perry (1987) found that
of the fungus and the IAA concentration of the soil from established plantations greatly increased
roots, indicating a possible relationship between mycorrhizal develop- ment in sites with repeated
relative capacity for IAA or ethylene production plantation failure. In con- trast, Rose et al. (1983)
and mycorrhizal forma- tion” (Scagel and found that litter could inhibit mycorrhizal
Linderman 1998, p. 746). These results support the formation. MacFall (1994) summarized literature
hypothesis that mycorrhizal fungi can stimulate indicating that ”ectomycorrhizae have the
increases in root IAA that can affect the growth of capability of altering the rhizosphere biogeochem-
roots and shoots after transplanting. Dosskey et al. istry, and of creating mineralization patterns
(1990) found that Rhizopogon vini- color, but not which differ from those of bulk soil” (p. 217).
Hebeloma crustuliniforme or Laccaria lac- cata, In his review of mycorrhizae in ecosystems
caused a significant increase in photosynthesis of world- wide, Read (1991, p. 379) noted that
Douglas-fir seedlings, and that the probable cause ectomycorrhizae are found on soils with a leaf
was the increased photosynthetic sink of extensive litter accumulation with a relatively low
fungal growth. In a second paper (1991), they noted carbon:nitrogen ratio, which forms a layer of
the same effect of R. vinicolor under conditions of acidic, organically enriched material on the soil
drying soil. surface characterized by such low rates of N
Simmard et al. (1997) elegantly demonstrated that mineralization that this element is frequently
the hyphae of mycorrhizas could translocate photo- growth limiting. The layer may be a mor, a
medium mull, or a mull when over base-rich
substrates. Read (1991) noted that many
mycorrchizal fungi have the
Chapter 11. Mycorrhizae 243
relation-
potential to mobilize organic nitrogen and, hence,
are more important ecologically than previously
thought. Read (1991) observed the proliferation of
absorptive hyphae in areas of local nutrient
concen- tration and described examples, such
“fungal wefts associated with the mycorrhizal
roots proliferating in the decomposition horizons
of coniferous forest soils and in the mycelial mats
formed by Hysterangium and related species
underneath the organic horizons of Douglas-fir
forests” (p. 382). Read (1991, p. 382) reported
that “it has been estimated that the myce- lium of
H. crassa can occupy 9.6% of the A horizon of a
forest soil to a depth of 10 cm. These mats
are known to be sites of elevated enzyme and re-
spiratory activity.”
According to MacFall (1994), “fungal mats
formed by Hysterangium setchellii (Fisher) in
association with Douglas-fir may colonize up to
27% of the forest floor and account for 45-55% of
total soil biomass” (p. 227). Read (1991)
suggested that “late” stage fungi are likely to
colonize seedlings and, hence, to integrate them
into a network of absorptive mycor- rhizal fungi.
Such integration can compensate for poorly
formed root systems and result in greater seedling
growth.

Fire
The numerous references on the effect of fire on
mycorrhizae are erratic and probably reflect the
intensity of a given fire, with fewer mycorrhizal
fungi in severly burned sites.

Cost-Benefits of Mycorrhizae
We have previously noted that mycorrhizae are
generally essential to the survival of the host
plants (including Douglas-fir). However, it may
be of inter- est to examine the cost, in terms of
plant resources, of this relationship. Fitter (1991,
p. 350) noted that about 10% of the carbon
translocated to the root goes to mycorrhizae.
However, this may not be a drain on plant
resources if it stimulates a higher rate of
photosynthesis. Further, photosynthesis rates are
frequently limited by phosphorus, an element
largely taken up by the fungi. Fitter suggested that
the rate of P uptake may very well be correlated
with productivity, although noting that such a
Which fungi stimulate which particular tree
ship had been found only in carefully species under which specific abiotic and biotic
controlled laboratory trials. Studies that conditions remains an open question, however.
attempted to relate mycorrhizae and yield Most biotic and abiotic parameters are not
resulted in erratic data. Fitter concluded that
more data describing uptake by roots and
mycorrhizae under field conditions are
needed. St. John and Coleman (1983, p. 1011)
cited numerous papers suggesting that a
mean of about 50% of the photosynthate is
translocated to fungi.

Conclusion
We have cited a number of papers that
discussed the relation between mycorrhizal
fungi and Douglas-fir, although this is not an
extensive review. Much is summarized in the
following from MacFall (1994):
It can be concluded that ectomycorrhizae play a
sig- nificant role in soil biogeochemistry and soil
structure. Greater selective nutrient uptake from
the increased absorptive surface area provided by
the mycelial network has been shown with H.
arenosa and other ectomycor- rhizal associates.
Although not clearly demonstrated, efficiency of
uptake is likely to also be increased with
mycorrhizae. Mechanisms for accelerated,
biologically- mediated weathering of minerals
and organic materials through the production of
enzymes, organic acids, and siderophores are
present in many ectomycorrhizae. Higher rates of
carbon and nutrient mineralization have been
observed within fungal mats compared to non-
mat soils, suggesting the potential for a similar
role in more diffuse hyphal structures.
Significant water uptake and transport may also
be accomplished through mycorrhi- zae, but is
likely to differ between mycorrhizal types.
Clearly these symbiotic associations have the
potential to alter the soil chemistry of the
mycorrhizosphere, and as our understanding of
their physiological processes emerges, we can
better develop a model for their role in nutrient
mobilization and cycling at the ecosystem level.
(MacFall 1994, p. 232)

As Amaranthus (1994) and later Steinfeld


et al. (2003) observed, more information is
needed “on the ability of specific mycorrhizal
fungi to establish at the nursery and improve
seedling performance in the outplanted
environment,” particularly given that
“fertilizing and irrigating practices in seedling
production nurseries are very different than
field conditions at harsh outplanting sites”
(Steinfeld et al. 2003, p. 197). The importance
of the mycorrhizal habit in the establishment
and survival of forest trees is not in doubt.
244 Douglas-fir: The Genus Pseudotsuga
studied rigorously enough to allow coherent as the hosts and the genus Pisolithus as the
predic- tions across a wide range of conditions. inoculated mycorrizal fungus. Additional attention
Considerable research on the effects of needs to be focused on native plants and host-
ectomycorrhizal inoculation on the outplanting specific mycor- rhizal fungi. Although the
performance of forest seedlings continues. applicability of the results is not as wide ranging,
Unfortunately, most of the work is similar to the potential for outplanting improvement may
previous work using species of the family well be higher.
Pinaceae
12. Adverse Abiotic Factors
Richard K. Hermann

B
oth abiotic and biotic factors can have an the destruction of experimental plantings of
ad- verse effect on the growth and survival Douglas- fir by an unusually severe early frost on
of Douglas-fir. the plateau of central North Island.
Frost Identification of frost injury
Observations on frost damage and research on The damage inflicted on tree nurseries by the cold
frost hardiness of Douglas-fir have focused spell of December 1972 provided a drastic
primarily on the coastal variety. That focus example of the consequences of such a climatic
reflects the fact that coastal Douglas-fir is event (Hermann 1974). Although exact figures are
considerably less frost resistant than the interior not available, losses for all the nurseries in Oregon
variety (Sakai and Weiser 1973), and the and Washington probably amounted to several
predominant role of coastal Douglas-fir in the million seedlings. That event emphasized the need
areas of introduction outside its natural distri- for the correct identification of frost injury to
bution. The susceptibility of the coastal variety to avoid compounding losses in the nursery by
frost injury became a matter of concern as early as preventing outplanting or injured stock, with its
the last quarter of the 19th century in German risk of lost planting invest- ments and future wood
trials of Douglas-fir (Danckelmann 1884). production. Although frost injury is generally not
Frost leads perhaps more often than any other as serious in older trees as it is in seedlings,
abiotic factor to injury of Douglas-fir. Damage identifying the type of injury may help in deciding
can become particularly severe with the influx of on the need for salvage measures.
abnor- mally cold air masses. Such climatic
events, with their ensuing damages, have been Needles
recorded for the Pacific Northwest in November Injury to needles is probably the most common
1955 (Duffield 1956) and December 1972-January and usually the most easily recognizable kind of
1973 (Hermann 1977). Douglas-fir grown outside frost injury. Injured foliage tends to lose its
its natural range has also experienced particularly normal green color from 24 to 72 hours after a
injurious cold spells. Extremely low temperatures return to above-freezing temperatures; its color
in October 1908 (Abele 1909); the winter of changes to a reddish-brown, sometimes preceded
1928/1929, when temperatures dropped to a low by a purplish or dull gray hue (Hermann and Zaerr
of −45°C (Kahl 1930); October 1955 and February 1973). Color change however, may not occur for
1956 (Jahnel and Watzlawick 1957); and in the fall several days if temperatures remain below
and winter of 1978/1979 (Jestaedt 1980), resulted freezing. Nevertheless, the possibility exists to
in widespread injury to coastal Douglas-fir in identify damage immediately. The freezing of
German nurseries and plantations. The winters of plant cells ruptures their membranes and alters
1923/1924, 1928/1929, and 1946/1947 caused electric properties of membranes. Changes in
much damage to saplings and pole-sized stands of electric impedance can be determined with special
coastal Douglas-fir in Denmark (Thulin 1949). A equipment; thus, damaged needle tissue can be
report from New Zealand (Director of Forestry de- tected minutes after it has thawed out (Zaerr
1943) mentions 1972).
245
246 Douglas-fir: The Genus Pseudotsuga

Buds responding to chilling, because the message for that


Frost injury to buds is not immediately apparent in initiating ac- climation (Hermann 1974). If needles
from their external appearance. Injured buds are removed or killed in early fall, buds appear to be
begin to look dried out or start to shrivel 4 to 6 incapable of
weeks after frost injury. If buds are injured,
however, damage can be determined by slicing
them open shortly after temperatures are again
above freezing. Damaged tis- sue inside the bud,
most commonly the leaf primor- dium, develops a
light-to-dark brown discoloration (Hermann and
Zaerr 1973).
Trunk
Frost injury to trunks of seedlings, saplings, and
pole-sized trees may show external signs, such as
lesions or shriveled bark. Depending on the
severity of injury, various degrees of browning by
cambium and phloem can appear after some bark
is removed. Another indication of damage to the
trunk is pro- gressive defoliation. Needles may be
shed even if they themselves have not been killed
by frost. That phenomenon appears to be
associated with injury to the needle traces and
cortical tissues of the trunk. Injury to the boles of
mature trees is rare and usu- ally is not visible.
Roots
Injured roots show brown or almost black discol-
oration when bark is stripped away. Injured bark
becomes mushy and can easily be pulled off. At
this stage, however, freezing injury can easily be
mis- taken for symptoms of fungal diseases
(Hermann 1990).

Consequences of frost injury


Frost injury is seldom severe enough to kill the
entire tree immediately, but it may weaken the
tree enough for it to die if it is stressed.
Needles
If seedlings lose most or all of their frost-damaged
foliage without injury to buds and stems, the loss
is seldom lethal. But it can still have serious
conse- quences. Artificial defoliation of 2-year-old
seedlings in August, November, and February to
investigate the importance of needles in the
dormancy cycle of Douglas-fir indicated their role
response appears to be routed through the seasons behind that of trees whose terminal bud
foliage. The consequences are increased had not been damaged. But by the end of the
susceptibility to frost injury, delayed bud third growing season, the height increment of
burst, and shoot growth in spring. If needles seedlings whose initial terminal bud had been
are lost after hardiness has been acquired,
bud burst and shoot development will be
normal.
Needle loss by saplings, even when severe
is not necessarily lethal. The development of
6-year-old trees in Christmas-tree plantations
injured by the cold spells in December 1972
and January 1973 was followed for 3 years.
Large losses of needles without severe injury
to the trunk had not resulted in either
immediate or delayed death of trees. Of a
sample of 2,000 trees, less than 5% had
suffered a needle loss exceeding 80%.
Severity of injury to needles did not always
indicate injury to other tissues, al- though
extensive damage to trunk tissues was most
frequently found in trees with severe foliar
injury. The trees were an economic loss,
however, because they had become
unmarketable (Hermann 1977). Zieger et al.
(1958) reported that the survey of frost
damage to coastal Douglas-fir in eastern
Germany by the February 1956 freeze showed
that only 6% of the frost-injured trees never
recovered. They were mostly saplings and
suppressed pole-sized and mature trees that
had lost between 70% and 80% of their
foliage.

Buds
The main effect of frost injury to the
terminal bud of 1- and 2-year-old seedlings
appears to be a tem- porary growth
reduction lasting through two or three
growing seasons, unless combined with se-
vere damage to the trunk and foliage. In a
study by Edgren (1970), who followed the
development of 2-0 Douglas-fir seedlings
injured by an early frost in September 1965
at the Wind River Nursery near Carson,
Washington, through three growing seasons,
95% of 500 seedlings outplanted in the field
failed to burst terminal buds in spring 1966.
A shoot from a lateral bud took over as the
new leader in these trees. The height growth
of damaged trees lagged for two growing
Chapter 12. Adverse Abiotic Factors 247
the tree. Branches up to a height of 50 cm height had
killed had caught up with and even surpassed the 65% of dead buds com- pared to 29% on branches at
height increment of trees whose terminal bud had heights of 151 to 200 cm.
escaped damage. A similar phenomenon,
illustrating the remarkable regenerative ability of
Douglas-fir to replace a damaged terminal bud or
shoot, was reported from Germany (Mörmann
1956a). A late frost in May 1953 had killed the
terminal shoots of 4- to 6-year-old trees in
plantations of coastal Douglas-fir. Not only did
these trees quickly form new leaders from a
lateral bud, but many of them displayed unusually
large increments in the next growing season.
Buds are especially prone to injury by late
frosts because deacclimation has usually ended
by that time. Seedlings and saplings growing in
valleys and depressions frequently suffer from
repeated frost damage in the spring. The frost
damage delays or even prevents the growth of
trees beyond the height of the layer of air
subject to freezing temperatures. The trees tend
to acquire a shrub-like appearance that is
sometimes mistakenly attributed to wildlife
browsing. Damage inflicted on buds by a late
frost is clearly illustrated by an event in May
1967 in the Soltau forest district in northern
Germany (Rack 1974). The injured trees were
coastal Douglas-firs, mostly 2- to 5-year-old. A
first inspection of the east- west aligned rows of
trees after the freeze hinted they had suffered
more damage on their south- than north-facing
side. A count of killed buds on 15 trees
confirmed that damage was indeed greater on
their south than on the north side. The tally
showed an average of 42 dead buds (range, 1–
100) on the south and 31 dead buds (range 0–
83) on the north side. Rack (1974) attributed the
difference in dam- age to an earlier bud break
on the south side and, therefore, buds on that
side were more damaged by a late frost. Rack
came to the conclusion because the
developmental stages of buds on the two sides
showed corresponding differences.
The temperature gradient in the air layer near
the
ground also contributed to differences in the
amount of damage (Rack 1974). A count of dead
buds on 18 trees at five 50 cm-intervals from
ground level to a height of 200 cm showed a
decrease of killed buds with increasing height on
nearly equal buds or by a whorl of short shoots,
In one of the Christmas tree plantations in each bearing a complete set of buds. Van der
Oregon’s Willamette Valley, the December Kamp and Worrall (1990) concluded that the
1972 freeze killed the terminal buds in 69 major long- term damage from this type of bud
trees out of 1,000 sampled (Hermann 1977). injury probably
Of the 1,000 trees, 530 had sustained injury to
lateral buds on the trunk. The number of
injured buds on the trunk was generally five
to six of those checked on each tree.
Fortunately, injury to the terminal bud does
not pose a prob- lem to Christmas tree
growers because the leader is usually pruned
away. Massive injury to lateral buds would be
more serious because of the gaps in the
crown. But as this survey suggested, massive
injury to lateral buds is probably uncommon
even during a severe freeze in mid-winter.
Frost damage to lateral buds in spring,
especially in consecutive years, is more likely
to render trees unmarketable. Some of the
trees sampled in spring 1973 apparently had
suffered bud injuries in previous years and, in
these trees, the additional injuries from the
freeze of December 1972 were sufficient to
make them un- marketable. Plantations and
young natural stands of interior Douglas-fir in
the central interior of British Columbia
showed widespread injury or death of buds in
parts of trees above the winter snow line in
spring of 1989 (Van der Kamp and Worrall
1990). That damage was apparently the result
of a rapid drop of temperatures from above
freezing to −30°C, when a mass of arctic air
moved into the region on January 30, 1989.
That event was preceded by four unusually
warm months, with temperatures of 1.0°C to
4.7°C above normal. Damage became visible
in mid-June, when most buds above 1 m (the
putative mid-winter snow line) had failed to
flush. Buds below that line had flushed
normally. Inspections in severely affected 10-
to 15-year-old plantations east of Williams
Lake, British Columbia (52°20’N, 121°30’W)
indicated that, of the buds above the snow
line, fewer than 5% had been killed, but
between 50% and 90% were severely injured.
In most of the damaged buds, the central
meristematic dome had been killed, but injury
was rare to the basal stem seg- ments bearing
the bud scales. These injuries resulted in the
replacement of the terminal bud by a whorl of
248 Douglas-fir: The Genus Pseudotsuga
increased the frequency of multiple leaders, 1966, over most of Vancouver Island injured cones 35
affecting adversely the form of the main bole. to
Reproductive buds
Observations in the field and freezing tests have
shown that reproductive buds appear to be the
least cold-hardy organs. Frosts during flowering
periods have caused heavy seed losses in
Douglas-fir stands and seed orchards in the native
range of this spe- cies and in areas of introduction.
The magnitude of such losses was illustrated by
the damage done by frosts in winter 1972/73,
which killed 77% of all female buds in a
Weyerhaeuser seed orchard in Oregon (Timmis
1977). The research prompted by this event
indicated that the hardiness of female buds
increased to a maximum by mid-December. A
50% kill (LT50) of female buds in freezing tests
ranged from −19°C and −23°C from mid-
December to early March. Hardiness then
decreased at 1.6°C per week to a LT50 of −6.5°C;
by the time the first external morphological
changes in buds were visible, hardiness had nearly
reached its minimum. After bud swelling,
significant differences in hardiness were not
found between female flowers at different stages
of their development. LT50 averaged −4.5°C
throughout flowering for the five clones used in
this investigation (Timmis 1977).
Temperatures leading to frost kill of female
buds on older trees during flowering apparently
are in the same range as those for female buds on
younger trees. A frost with temperatures between
−2°C and
-3°C in spring of 1953 completely destroyed
flowers in 60- to 70-year-old coastal Douglas-firs
in eastern Germany (Krauss 1955). Male and
vegetative buds are apparently slightly hardier
than female buds. Timmis (1977) found male
buds to be 1°C or 2°C and vegetative buds 2°C or
3°C hardier than female buds, especially from
March to May. The January 1989 freeze in central
interior British Columbia had only minor effect on
vegetative buds of mature in- terior Douglas-fir,
but apparently killed all of their reproductive
buds, both male and female. The 1989 Douglas-fir
cone crop was a complete failure in that region
(Van der Kamp and Worrall 1990).
Frost damage to immature cones is probably
rare. A severe frost in the night of May 26–27,
45 mm in length at the Gordon River clone destroying fungi and the prevalence of shake in
bank. Ovuliferous scales were discolored and outwardly sound-looking trees would result, at
the central axis of the cones showed complete best, in poor quality wood from the butt log and
necrosis. Bract scales, however, were still that pulp would be the most likely end product.
green (Orr-Ewing 1966a). Another such event
was recorded in southwest- ern Germany
where a late frost on May 23, 1953, destroyed
immature cones on 60-year-old coastal
Douglas-firs (Mörmann 1956a).
Stem tissue
Injury to stem tissues, especially the phloem
and cambium, are generally more critical than
injury to either buds or needles. A seedling
usually will die if the injury extends over
more than half the circum- ference of the
stem. Such injury has essentially the same
effect as girdling the stem. Seedlings injured
in this manner often do not develop external
signs of injury and may even commence to
grow new shoots before dying (Hermann
1974).
The long-term consequences of frost
damage to the trunk of older trees have been
shown particularly well in the aftermath of the
November 1955 cold wave in the Pacific
Northwest. Severe frost damage sustained by
an 11-year-old Douglas fir plantation near
Elma, Washington, resulted in top kill
through- out the stand. The stand’s average
height of 4.3 m before the freeze dropped to
2.4 m immediately after that event. Five years
later, the height of injured trees that had
developed new leaders averaged 5.5 m versus
6.4 m for leaders of uninjured trees. Injured
trees that had failed to form a definite new
leader averaged only 4.0 m in height (Wiley
1960).
A follow-up study (Shea 1962) of injured
trees in the Elma plantation at age 17
indicated invasion by several species of
wood-destroying fungi, as well as other forms
of damage. Frequently, damage to the
cambium had resulted in the formation of a
partial or complete frost ring. Shake had
developed in some trees and had separated
wood formed before the freeze from that
formed afterwards in portions of the trunk.
Partial separation of the 1955/1956 growth
rings had occurred in other trees. Shea
concluded that the presence of wood
Chapter 12. Adverse Abiotic Factors 249
moderately damaged by the November 1955 freeze in
A case in which serious frost injury to trunks the Pacific
by the November 1955 freeze that had not become
im- mediately evident was described by Childs
(1961). In 1957, hundreds of small groups of
trees, appar- ently scattered randomly, showed red
crown in an extensive well-stocked stand of
Douglas-fir in northwestern Oregon. Trees ranged
in age from 25 to 30 years, averaging 20 cm in
dbh. In 1958, many of these trees were dead.
Examination revealed callus layers, as wide as an
annual growth ring, completely encircling the
boles between 2.45 and 3.56 m above their bases.
Smaller banks of callus tissue, about one-half to
two-thirds as wide as the preceding an- nual ring,
were also found slightly higher on their trunks.
Many of the still-living trees, immediately
adjacent to the groups of killed ones, displayed
cracked bark about 0.6 to 2.7 m above the base as
a result of callus formation where small areas of
cam- bium had been killed. Conspicuous frost
rings were present between the 1955 and 1956
annual rings. In most of the trees with partially
killed cambium in the basal part of the bole, the
upper 1.2 to 1.5 m of the crown had died.
Although information about further development
is not available, most of this stand has probably
become a total loss because of the severity of
injuries sustained.
Johnson (1971) followed the development of
trees injured by the November 1955 freeze in a
17-year- old plantation on Vancouver Island. He
examined the frost lesions on severely injured
trees at 3-year intervals. Three years after the
initial injury, the frost lesions had exposed
sapwood. Callus tissue had formed over 50% of
the lesions within 6 years of the injury and over
70% after 11 years. Girdling by either a single
large lesion or multiple lesions fre- quently
resulted in the death of the leader. In all these
cases, a lateral shoot had assumed dominance, and
little evidence of damage was found after 11
years. Johnson’s statement that “this study
revealed that even with severe frost injury
Douglas-fir of this age recovers rapidly with little
lasting effect” is perhaps too much of a
generalization in view of the findings of Childs
(1962) and (Shea 1962).
A study by Reukema (1964) demonstrated the
effects of frost injury on the radial growth of 50-
year- old Douglas-fir that appeared to be only
teractions of environmental, genetic, and
Northwest. He analyzed the radial growth physiologi- cal factors. The first stage of cold
of 14 codominant trees from a site III stand acclimation usually begins in early fall with the
in western Washington for the years from onset of cool nights and shortening photoperiods
1950 to 1959. The trees, averaging 31.5 cm (Irgens-Møller 1957, Van den Driessche 1969a).
in dbh and 30.5 m in height, had been felled When night temperatures drop
and dissected for growth analysis.
Fluctuations in the amounts of annual
growth during the 10-year-period (1950–
1959) were com- mon and were apparently
related to variations of temperature and
precipitation. In none of the other 9 years,
however, were growth reductions as great as
in 1956. Ratios of 1956 growth to 1954–1955
average growth in internodes 20 to 40,
roughly from base of live crown to breast
height, ranged from 10% to 71%, averaging
57%. The range in growth reduction reflects
the fact that trees that had the fastest growth
before the freeze suffered the greatest
reduction of growth. In general, growth was
more curtailed in the lower part of the trunk
than near the base of the crown, but even in
the crown, stem growth was re- duced
substantially. Radial growth began to improve
in 1957, but it had still not reached the growth
rate
of the years 1954 and 1955 by 1959.
Roots
Lethal injury to roots beyond the seedling
stage is probably rare. Soils seldom reach
temperatures low enough to kill the roots of
older trees. Direct kill of roots of seedlings
in nurseries or newly established plantations
appears also to be rare, which may be
inferred from a study in a Scottish nursery
(Cannell et al. 1990). The study
demonstrated that hardiness levels of roots
of 2-year-old seedlings were already well
above lethal soil temperatures in October.
We are aware of only one report (Soljanic
1968) attrib- uting the mortality of Douglas-
fir seedlings in a plantation to frost kill of
roots.

Cold acclimation and deacclimation


Douglas-fir, like other coniferous evergreens,
under- goes seasonal changes in frost
hardiness. The pro- cesses that lead to the
development and loss of cold hardiness are
governed by complex and varying in-
250 Douglas-fir: The Genus Pseudotsuga
to near freezing, they trigger the beginning of between and within populations of Douglas-fir. As
the rapid (second) hardening stage that leads to Larson (1978) and Anekonda et al. (2000) pointed
the peak of hardiness in mid-winter (Lavender et out, “in- cidence of natural frost injury is typically
al. 1968, Van den Driessche 1969a). During winter sporadic
rest, Douglas-fir cannot be induced to grow until
its chilling require- ments are met (McCreary et
al. 1989). Then, growth resumes under favorable
conditions. Deacclimation is induced in late
winter or early spring by rising temperatures
(Van den Driessche 1969a, Schuch et al. 1989a).
But experiments by Worrall and Timmis (1974,
1975), with 2-year-old seedlings of coastal and
interior Douglas-fir, indicated that only the
initial dehardening occurs in response to
warming temperatures. Their findings suggested
that final loss of hardiness depended on growth-
promoting hormones from expanding shoots,
indicative of a two-stage dehardening process,
presumably pre- venting premature loss of
hardiness.
Alden (1971) found that acclimation and
deaccli-
mation is not simultaneous in all tissues of
Douglas- fir. Working with cut terminal twigs of
10-year-old Douglas-fir from a Willamette Valley
seed source, he demonstrated, seasonal variation
in cold hardiness among tissues of the stem,
needles, and terminal and lateral buds. For
instance, injury in early winter to the most
susceptible tissues in hardened twigs, such as
transfusion tissues and interfascicular parenchyma
of the bud trace, began at −15.1°C with slow cool-
ing rates, and all cells in these tissues were killed
at
−30°C. By contrast, the more resistant bud scales
and cortex could be cooled to −30°C before injury
became evident. Furthermore, the relative
susceptibility of some tissues to freezing injury
changed during the development and loss of
hardiness. The pith, for example, was more
susceptible to injury than were other tissues of the
stem in fall, but more resistant in spring. Timmis
(1976), Aitken and Adams (1996), and Rose and
Haase (2002) also reported relative seasonal
responses to freezing stress among tissues.
Assessment of cold hardiness
The assessment of cold hardiness after damaging
frost events in the field is, in general, poorly
suited for studying variations in cold hardiness
across field test sites and over time, leading to Time of bud set and bud burst has been used as
un- even testing and poor statistical an indirect assessment of cold hardiness.
precision.” To avoid these problems, Genotypes having the earliest bud set can be
investigators have turned to artificial freeze expected to be the least susceptible to injury by a
testing, where the temperature of freezing fall frost, and those late flushing to be the
whole plants of their detached parts can be hardiest when exposed to a
strictly controlled and uniformly applied.
Frost injury can be assessed quantitatively by
measuring freeze- induced electrolyte leakage
(Burr et al. 1990) and chlorophyll
fluorescence (Rose and Haase 2002) of
freeze-tested tissue, or subjectively by visual
scor- ing. Cold hardiness is usually expressed
either as the temperature that inflicts lethal
damage to 50% of tested tissues or simply as
the percentage of damage at one or more
temperatures.
A simple method for assessing cold
hardiness using visual scoring of frost injury
was developed at the Forest Research
Laboratory of Oregon State University
(Anekonda et al. 2000). Frozen tissue is
allowed to develop damage symptoms for
several hours after freezing and then is
scored into damage classes. The method has
the advantage that large numbers of samples
can be efficiently processed at one time
because the visual assessment of injury is
quick and easy. Although injury is assessed
less accurately than with electrolytic
conductivity or chlorophyll fluorescence,
visual damage scores have been found to be
strongly correlated (r2 0.90), with injury
assessed quantitatively by freeze-induced
electrolyte leakage (Shortt et al. 1996) and
chlorophyll fluorescence (Rose and Haase
2002).
To what extent the results of artificial freeze-
testing reflect cold hardiness under actual
field con- ditions is not clear. Aitken and
Adams (1997) and O’Neill (1999), however,
in their studies of fall and spring cold
hardiness of breeding populations of coastal
Douglas-fir found moderate to strong es-
timated genetic correlations between visual
injury scores after artificial freezing and
natural frost events (mean = 0.82, range =
0.47–1.00). Their findings sug- gest that
artificial freeze-testing can provide reliable
predictions of the relative hardiness of
families to both fall and spring frosts.
Chapter 12. Adverse Abiotic Factors 251

spring frost. Anekonda et al. (2000) stated, by cit- in fall, winter, and spring cold hardiness of both
ing several references, that these expectations are the coastal and interior varieties of Douglas-fir.
generally met in coastal Douglas-fir. Moreover, significant variation has been found
not only between but also within populations. But
Age as Aitken and Adams (1996) wrote, “Genetic
Douglas-fir is most susceptible to frost injury as variation in cold hardiness of Douglas-fir appears
a seedling and becomes less likely to suffer dam- to result mostly from variation in timing of
age with advancing age. Experiments by Timmis acclimation and deacclimation rather than
and Worrall (1975) demonstrated that seedlings of variation in absolute levels of cold hardiness
both coastal and interior Douglas-fir were unable achieved.”
to develop freezing tolerance in the initial weeks
after germination and hence were extremely
Fall cold hardiness
vulnerable to frost damage. But even as their An unintentional mass selection made in 1883 by
ability to acclimate increases gradually with age, the forerunner of the Saxon Forest Experiment
seedlings still remain very vulnerable, which is Station (Bellmann and Schönbach 1964) led, 65
attributable to their small size, their tendency to years later, to the initiation of a study that
continue growing into late summer or fall provided, perhaps for the first time, proof of
(Campbell and Sorensen 1973), and their genetic variation in fall cold hardiness within a
proximity to frost layers near the ground. That population of coastal Douglas-fir. Seed, probably
vulnerability has been particularly well of western Washington origin, pro- vided by J.
documented by reports of extensive frost damage Booth was sown in a nursery in spring 1882.
to seedlings in forest nurseries. Examples in the Many of the seedlings suffered frost damage in
Pacific Northwest are the losses caused at the early winter of 1882.
Wind River Nursery of the USA Forest Service by Frost-damaged and uninjured seedlings were
a fall frost in 1916 (Munger and Morris 1936), in transplanted separately in spring 1883 and were
early September 1965 (Edgren 1970), and by a later planted in the Tharandt Forest, with the
late spring frost on May 31, 1933 (Munger and dam- aged plants in compartment 29 and the
Morris 1936). uninjured in compartment 50 (Schönbach
Although trees become less vulnerable to frost 1953). Cones were harvested in 1948 from 10
injury as they mature, they may suffer damage in open-pollinated, then 65-year-old trees in each
un- usually severe freezes such as the one in of the two compartments, 29 and 50. Seeds from
November 1956 in the Pacific Northwest or the each of the 20 trees were sown separately in
extremely cold winter of 1928/29 in Europe. An spring 1949. A night frost in October 1949
illustration of de- creasing frost damage with caused injury to progeny from the mother trees in
increasing age is pro- vided by a survey after the compartment 29, but little or none to progeny
February 1956 freeze in eastern Germany (Zieger from mother trees in compartment 50.
et al. 1958). Results of the survey listed by 20- Seedlings were outplanted as 3-year-old
year age classes indicated that most of the damage trans- plants in two different locations, one
was in the youngest age class (Table 12.1). The under shelter and the other without shelter.
frost, which lasted through all of February 1956 Eight-year survival
as a result of the influx of arc- tic air masses, was
preceded by an unusually mild Table 12.1 Percentage of trees damaged by the February
1956 freeze in eastern Germany, listed by age class (from
Zieger et al. 1958).
December and January. This particular sequence of
climatic events may have led to early Genetics of cold hardiness
deacclimation of trees, and thus compounded the
Observations of frost injury in the field and in labo-
damage caused by the February freeze (Zieger et
ratory experiments have shown genetic variation
al. 1958).
Age class Injured trees %
1 – 20 37
21 – 40 20
41 – 60 14
61 – 80 13
81 – 100 10
101 – 120 6
252 Douglas-fir: The Genus Pseudotsuga
of progeny from mother trees in compartment seasons. Hardiness of the needles was nearly the
29 and 50 in the plantation without shelter same from May to September 1958 for all
(Table 12.2) showed a striking difference families with an LT50 of about −8°C. Hardening
between the two. The low survival of progeny of family
from mother trees in com- partment 29 was 20 saplings began already in September, ahead of
almost entirely the result of losses from fall and the other families by nearly a month (Figure
winter frost injury (Schönbach 1958). The 12.1). Hardiness increased to a LT50 of −21°C by
difference in frost hardiness between the par- the end of January 1959 for families 20 and 23,
ent trees in compartments 29 and 50 was further with a LT50 of about −16°C for families 3 and 4 to
demonstrated by Schönbach (1958, pp. 329, begin decreasing again from February to May.
355) with cuttings from these trees. Difference Scheumann’s findings that differences in cold
between vegetatively propagated progeny were hardiness among families continued to find
less pro- nounced, however, than those between expression with increasing age of trees
sexually propagated progeny. corroborated Schönbach’s (1959) results.
Scheumann (1962), in a complementary study,
used freezing tests of needles from 10-year-old
Quantitative genetics
prog- eny of two parent trees (No. 20 and 23) in Coastal Douglas-fir
com- partment 50 and two parent trees (No. 3 and
A comprehensive picture of the quantitative ge-
4) in compartment 29 to investigate the seasonal
netics of fall, winter, and spring cold hardiness in
course of their frost hardiness. Families 20 and 23
coastal Douglas-fir has emerged from a series of
con- tained 96% and 76%, respectively, of the
investigations supported by the Pacific Northwest
originally outplanted trees. In families 3 and 4,
Tree Improvement Cooperative. The materials
only 32% and 19% of the originally outplanted
used by the investigators represent 291 families in
trees had not been killed by frost. An LT50 was
five breeding populations from British Columbia
used as a measure of (1), Washington (2), and Oregon (2). Artificial
the relative frost hardiness over the course of
the freez- ing of cut terminal shoots from first-order
lateral branches of 4- to 7-year-old saplings in the
10
laboratory and subsequent visual scoring of
damage to stem, needle, and bud tissues were used
-5 in these studies to assess cold hardiness.
Fall cold hardiness. Studies of cold hardiness
Temperature (°C)

-10
in the fall were conducted with four breeding
-15 popula- tions, two in western Washington (Aitken
20 Hardy et al. 1996) and two in western Oregon (Aitken
-20 23
3 Less hardy and Adams 1996). Investigations of cold
4
-25 hardiness in winter were added in the Oregon
study.
-30 The two studies demonstrated considerable
A M J J A S O N D J F MA M
with- in-population variation for fall cold
1958 1959 hardiness in both the Washington and Oregon
breeding populations.
Figure 12.1 Course of relative frost hardiness in four single tree Considerable variation among families was apparent
progenies from April 1958 to May 1959 (from Scheumann 1962).
for cold injury of stems, needles, and buds in early

Table 12.2 Percent survival of progeny from mother trees in compartments 50 and 29 in the first 8 years after outplanting (from Schönbach
1959).

Fall 1952 1953 1955 1956 1957 1958 1959


Compartment 50 77.3 74.7 73.4 72.9 72.6 72.0 72.0
Compartment 29 40.1 36.2 25.0 20.2 16.8 16.3 16.2
Chapter 12. Adverse Abiotic Factors 253
stronger genetic
fall to mid-fall, but differences were often small
or insignificant in late fall and mid-winter. The
genetic correlations for cold injury among these
tissues were generally not strong, indicating that
genes controlling hardening in the fall differ
somewhat for different tissues. Genetic
correlations between fall and winter cold
hardiness suggested that hardiness at these two
stages is apparently under separate genetic
control. Moreover, mid-winter cold hardi- ness
was shown to be weakly inherited. Therefore,
selection for cold hardiness in fall will probably
have little, if any, impact on hardiness to extreme
winter temperature. Because cold injury to coastal
Douglas-fir in mid-winter is unlikely to pose
much of a problem in the Pacific Northwest, the
region’s breeding programs seem to have little
need to con- sider midwinter hardiness as a
separate trait (Aiken and Adams 1996).
Spring cold hardiness. Spring frosts are common
in the natural range of Douglas-fir and in regions
of its introduction, and often cause considerable
damage to Douglas-fir in nurseries, young planta-
tions, and recent natural regeneration. Injury
caused by spring frost may occur before bud burst
when tissues deacclimate or after bud burst as
damage to newly flushed shoots. As stated by
Aitken and Adams (1997), bud phenology was
widely used to indicate spring cold hardiness
while knowledge remained incomplete about the
genetics of spring cold hardiness before bud break
and the degree of variation in the rate of
deacclimation among families. Aitken and Adams
(1997) began a study in 1993 to close that
knowledge gap. The material for the in-
vestigation were 7-year-old saplings from the
same two breeding zones—one in the Cascades,
the other in the Coast Ranger—and test sites for
the study of fall cold hardiness in western Oregon
(Aitken and Adams 1996). Samples consisted of
shoots cut in March and April 1993 and in April
1994, subjected to artificial freezing, and visually
scored for cold injury to stems, needles, and buds.
Additionally, in April and May of 1993, bud burst
was recorded bi-weekly on a single, marked
branch of each tree from which shoot samples had
been collected.
Results of the study indicated that cold
hardiness in spring is under considerably
1997). Seedlings were grown in raised nursery
control than is cold hardiness in the fall. beds and subjected to two soil moisture regimes,
Individual heritabilities for scores of spring one avoiding and the other creating moisture
cold injury aver- aged 0.76 in the coastal zone stress,
and 0.42 in the Cascade zone. Conversely,
heritability estimates for scores of cold injury
in mid-fall, with material from the same
breeding zones and test sites, were all under
0.40 and averaged 0.27. The authors surmised
that the high heritabilities for cold injury in
spring suggest that fewer genes may control
cold hardiness in spring than in fall. Although
heritabilities for cold injury in fall were low,
Anekonda et al. (2000) considered them to be
still large enough to permit good progress in
genetically improving cold hardiness by
selection and breeding.
Another finding related to understanding
the processes involved in the development
and loss of cold hardiness is that cold
hardiness of stems, needles, and buds is
strongly correlated in the spring, contrary to
weaker correlations in the fall (Aitken and
Adams 1996). Thus, the authors proposed that
different shoot tissues deacclimate apparently
in synchrony in response to the cumulative
effects of chilling and heat sum, whereas cold
acclimation rates and timing vary among
tissues. Genetic correlations between spring
cold injury scores for all tissues and date of
bud burst were strong and negative, with
genotypes that break buds early having high
cold injury scores for all tissues.
Seedlings versus saplings. The studies on
the genetics of cold hardiness of coastal
Douglas-fir by Aitken et al. (1996), Aitken
and Adams (1996, 1997), and Anekonda
(1998) used saplings (4-to 7-year-old trees)
because of their widespread and immediate
availability in established progeny tests.
These in- vestigations were followed by a
study (O’Neill 1999) of the cold hardiness of
seedlings in fall, winter, and spring after their
second growing season. Seedlings were raised
from seed of open-pollinated Douglas- fir.
The seed stemmed from 40 parent trees within
each of two breeding zones in western
Oregon, one in the Cascades and the other in
the Coast Range. They represented the same
breeding populations used in the
investigations by Aitken and Adams (1996,
254 Douglas-fir: The Genus Pseudotsuga
to assess how drought during the growing season of secondary lateral branches were collected for
affected cold hardiness among families. Samples two kinds of artificial freezing tests from five
consisted of shoots detached in September, randomly selected trees per family. Four branch
October, and November, and in March before the tips from each sample were assessed for cold
third grow- ing season. They were subjected to injury by the visual method and the remainder
artificial freezing and then were visually of the branch by the freeze-induced electrolyte
evaluated for cold injury to stems, needles, and leakage (FIEL) method. The results indicated the
buds. lack of a strong con- sistent relationship between
The freezing tests demonstrated that, at the cold hardiness and inbreeding in Douglas-fir.
seed- ling stage, significant genetic differentiation Although different tissues were used in the two
exists between the Coast Range and Cascades kinds of hardiness tests, over- all, trends in
populations for cold hardiness in the fall, but not damage were found to be similar. All the analyses
in the spring. Despite its significance, genetic of the tests demonstrated that trees within
differentiation be- tween the two breeding common ancestor by inbreeding groups accounted
populations was for the most part exceeded by for significant amounts of variation. The authors
genetic variation among families within each of surmised that the large family and within-family
the two populations. Cold hardiness of seedlings variability in hardiness may have reflected a
in fall and spring varied widely among families in limited sample size or be explained by the range
both breeding zones, indicating strong genetic of physi- ological factors that can influence cold
control, with mean estimates of individual hardiness. They concluded that “inbreeding does
heritabilities for cold injury of 0.37 in fall and not appear to have a significant deleterious effect
0.57 in spring. Variation among families and of the frost hardiness of Douglas-fir from the
estimates of heritability (h2 = 0.22) were weak for coastal breeding zone; and loss of productivity
cold hardi- ness in winter, however. Seedlings that due to frost damage in related progeny should
experienced moisture stress during the growing not be a major concern for tree breeders.” Most
season incurred significantly less cold injury in likely, that conclusion may
the fall than did those grown without moisture be valid for coastal Douglas-fir in general.
stress. Family rankings for cold hardiness in fall The investigations that focused on the genetic
and spring were nevertheless fairly consistent aspects of cold hardiness in coastal Douglas-fir
across the two soil moisture regimes. O’Neill’s have demonstrated considerable genetic variation
study allowed the evaluation of cold hardi- ness in in both fall and spring cold hardiness within
the fall and spring at both the seedling and sapling populations. The relevance of that variation for
stage in the two Oregon breeding popula- tions. practical appli- cations was expressed by
High genetic correlations and thus consistent Anekonda et al. (2000) as follows: “This variation
family rankings suggested that cold hardiness in can be exploited in tree improvement programs in
seedlings and saplings appears to be controlled two ways: (1) in selection and breeding to
largely by the same set of genes. maintain or enhance levels of cold hardiness of
improved strains and/or (2) in choosing specific
Inbreeding. Lacking knowledge as to how in-
families for planting in frost-prone sites.”
breeding affected frost hardiness of conifers led
Shortt et al. (1996) to design an experiment to Interior Douglas-fir
inves- tigate the effect of inbreeding on cold Douglas-fir in the Inland Northwest occupies a
hardiness of coastal Douglas-fir in the spring. For range of extremely heterogeneous environments
their study, they selected 19 families representing with frost-free period varying from 60 to 150 days
four inbreeding levels within five common (U.S. Department of Commerce 1968). The first
ancestor groups from an experimental plantation in- dication of genetic differentiation of cold
established in 1988 with one- year-old seedlings on tolerance in populations of Douglas-fir in the
Vancouver Island. Founder clones came from the northern Rocky Mountains came from a
low-elevation coastal breeding zone in British provenance study by Wright et al. (1971) and an
Columbia. Sections, 10-cm in length, investigation by Rehfeldt (1974a)
Chapter 12. Adverse Abiotic Factors 255
representing the other two provinces, eastern
of local differentiation of Douglas-fir populations Washington and northern Idaho on one hand and
in northern Idaho. Rehfeldt’s subsequent Montana on the
investigations on the cold hardiness of populations
of Douglas-fir from eastern Washington, northern
central Idaho, and western Montana were aimed at
gaining knowl- edge as to how “differentiation of
populations in cold hardiness may reflect
ecological adaptations that should be considered
in limiting seed transfer for reforestation”
(Rehfeldt 1979b). Rehfeldt (1978) studied growth
and cold hardiness of 90 families, 5 each within
18 populations of open-pollinated trees from five
warm and three cool habitat types in northeastern
Washington, northern Idaho, and western
Montana. Seedlings had been raised in two Idaho
forest nurseries, one (Coeur d’Alene) repre-
senting a relatively warm environment and the
other (Priest River Experimental Forest) a
relatively cool environment. At the end of the fifth
growing season, twigs cut from current growth in
mid-September and in late November were
subjected to freezing tests. Injury to leaves, buds,
and stems, was scored visually by judging the
degree of discoloration of tissues. The
temperature at which injury to each tissue was
first observed was used as a measure of cold
hardiness.
Initial analysis of the results indicated that
differ-
ences among populations in regard to growth, bud
phenology, and cold hardiness could be related to
habitat types. Much of the effect of habitat type
ap- peared to be attributable to the differing
performance of saplings from cool compared with
warm environ- ments. Further analysis suggested
that populations from relatively warm
environments could be divided into two
physiographic groups: (1) western Montana and
(2) eastern Washington and northern Idaho.
Populations represented in the study apparently
reflected adaptations to three different
environments, which Rehfeldt termed
“environmental provinces.” Populations from
relatively cool environments—re- gardless of
geographic origin—represent an “eco- logical”
province. Populations in this province are
characterized by slow rates of growth, early bud
set, and, as shown by the freezing tests, high
levels of cold hardiness. The populations
associated with injury to 50% of the twigs for
other, differed in these three traits, both from each sampling date (Figure 12.2). Low levels of
each other and also from the “ecological” hardiness developed after the first frost on
province. October 3. After that date, hardiness increased
In his study on the genetic differentiation moderately during a relatively warm period in late
of Douglas-fir populations in the northern
Rocky Mountains, Rehfeldt (1978) showed
that cold tol- erance was generally greater in
Douglas-fir from western Montana than in
Douglas-fir from Idaho and eastern
Washington. In a subsequent study of
variation of cold acclimation among
populations of Douglas-fir, Rehfeldt
(1979c) focused on north- ern Idaho and
eastern Washington. Tolerance to low
temperatures is necessary for the survival of
seedlings and saplings because freezing
tempera- tures can occur during any month
of the year, and temperatures reaching
−40°C are not rare in win- ter. The study
was conducted with 47 populations from
northern Idaho and eastern Washington, but
included two populations each from the
Okanogan Mountains in north-central
Washington and the Blue Mountains in
northeastern Oregon. Freezing tests to
determine levels of cold hardiness were
conducted on six dates between August and
December in the second growing of
seedlings raised for the study at Moscow,
Idaho. Twigs cut from the current growth of
seedlings representing each population were
used for the freezing tests.
Rehfeldt assessed cold tolerance by
regression analyses that were made for each
population ac- cording to a mathematical
model. He made separate analyses for each
population on data obtained be- fore and after
the first frost on October 3, that is for phases
one and two of cold acclimation. Tolerance of
each population to freezing during both
phases of cold acclimation was expressed as
the injury predicted from the mathematical
model when 50% of the twigs from all
populations exhibited injury. Differentiation
of populations in relation to latitude,
longitude, and habitat type was assessed by
mul- tiple regression analysis. Rehfeldt
demonstrated a close correlation of cold
acclimation with daily minimum temperatures
at Moscow by estimating the temperature
256 Douglas-fir: The Genus Pseudotsuga
and moist. Additionally, eight populations from
10 northern and central Idaho were included for the
5
freezing tests.
The freezing tests were carried out on terminal
0
shoots cut from 2-year-old seedlings in mid-Sep-
-5 tember, when first autumnal frosts can be
expected. Rehfeldt chose that single date because
Temperature (°C)

-10
“northern Idaho populations expressed greatest
-15
differentiation in cold hardiness after bud set but
-20 prior to the first fall frost and because previous
-25
test indicated that the ranking of populations
according to hardiness did not change at
-30
subsequent sampling dates.” One set of shoots
-35 from each population was frozen at the rate of
-40
5°C/h to one of five test temperatures between
−14°C and −18°C. Injury to each shoot was scored
15 1 15 1 15 1 15 1 by discoloration of needles. Differentiation of
Aug Sept Sept Oct Oct Nov Nov Dec
populations was assessed by an analysis of
variance of random effects on the percentage of
shoots of each
Figure 12.2 Frost hardiness and injury (from Rehfeldt). ranging from dry and warm to cool

October, but increased greatly after the first cold


wave in mid-November.
Rehfeldt’s models suggested that latitude and
elevation of the seed source controlled differentia-
tion of populations for hardiness, but only during
phase one of acclimation. Longitude and habitat
type had little effect. Of the four populations from
areas peripheral to the central area of the study,
the two from the Blue Mountains in eastern
Oregon hardened similarly to populations from
northern Idaho. The two populations from the
Okanogan Mountains in north-central Washington
showed a different pattern of hardening. They
apparently approached maximal hardiness already
in mid-November and, unlike populations from
farther east, failed to become still hardier in
response to late November’s cold wave. In the
next study, Rehfeldt (1982) investigated patterns
of genetic variation among Douglas-fir
populations from Montana west of the Continental
Divide. He excluded the extreme northwestern
por- tion of Montana from the study, which
presented patterns of environmental variation and
genetic dif- ferentiation similar to those of
northern Idaho. The 50 populations selected to
represent the ecological amplitude in the region
came from sites differing in elevation by as much
as 1,300 m, and representing habitat types
population exhibiting injury at each test temperature.
The freezing test revealed significant
differences in cold tolerance among the 58
populations. Although test temperatures
spanned only 5°C, percentage of injury to all
twigs ranged from 19% at −14°C to 75% at
−18°C. Across this range of temperatures,
mean injury to populations from western
Montana ranged from 16% to 69%.
In the third of the series of investigations
of ecological adaptations in Douglas-fir
populations, Rehfeldt (1983a) looked at
adaptive differentiation of populations from
central Idaho. For that study, he used 69
populations representing the geographic
distribution and ecological amplitude of the
species in central Idaho. Five-month-old
container seedlings raised at Moscow, Idaho
(46°05’N, 116°07’W) were transplanted into
four test environments. Two of those were at
Moscow, at an elevation of 700 m and with an
average frost-free period of 130 days. The
other two environments were at the Priest
River Experimental Forest (48°05’N,
117°00’W). One was at 670 m elevation with
an average frost-free period of 90 days; the
second was at 1,500 m elevation where
growing seasons are extremely short and
snow com- monly covers the site for 8
months, reaching a depth of about 300 cm.
The test environments markedly influenced
the phenology of seedlings. Those at the three
lower elevations burst buds in the last week of
April and nearly all had already set buds by
mid-
Chapter 12. Adverse Abiotic Factors 257
potential for frost damage and to aid in the selection
August. Budburst at 1,500 m did not occur until of genotypes
the first week of June and only 77% of seedlings
had set buds by mid-August. Thus, phenological
differ- ences resulted in injury from early but not
late frost to seedlings at the lower elevation sites,
and those at the high elevation site suffered injury
from late frost only as well as from impacts of
snow depth.
Results of the study covering central Idaho
(Rehfeldt 1983a), as well as those of the two
previ- ous studies (Rehfeldt 1979c, 1982)
demonstrated that populations from within
northern Idaho, western Montana, and central
Idaho displayed a different tolerance to early fall
frosts when acclimating in a common
environment. But the cold tolerance ex- hibited by
a particular population in one study was not
directly comparable to that of populations in other
studies for several reasons listed by Rehfeldt
(1986a). In the northern Idaho study, cold
tolerance was assessed throughout the period of
acclimation; in the other two studies, tolerance
was observed on a single date. In the central Idaho
study, conducted as a field test, injury from a
single night’s natural frost, representing a single
treatment, resulted in a mean injury of 8%. In the
northern Idaho and western Montana studies,
injury resulted from ar- tificial freezing of
detached twigs at several test temperatures
representing numerous treatments that produced a
range in mean injury from nearly zero to almost
100%.
In an update of the models built from these
three studies, Rehfeldt (1986a) developed a single
model describing genetic variation in cold
tolerance of Douglas-fir in the Inland Northwest.
In Rehfeldt’s words (1983a), results have shown
“ecological ad- aptations that differentiate
populations within these three regions based on a
network of traits that reflect adaptations to the
cold. Populations from severe environments
display greater cold hardiness but lower growth
potential than populations from mild
environments. Genetic differentiation therefore, is
strongly related to environmental gradients.”

Frost hardiness prediction models


Models to predict frost hardiness of Douglas-fir
have been built for both of its varieties. They
represent attempts to identify areas with the
According to the model, the risk of fall frost is
suitable for establishing regeneration on frost- highest at elevations above 500 m in the Cascades
prone sites. and the Olympic peninsula, but lower at
Interior Douglas-fir comparable elevations in the Coast Ranges. The
risk of spring frost, strongly associated with
The model made by Rehfeldt (1986a) to show
elevation, is higher
genetic variation in cold tolerance during
acclimation was based on studies of
populations belonging to the northern
subgroup of interior Douglas-fir within the
Inland Northwest. The model described
genetic variation in fall frost tolerance of
Douglas-fir along relatively steep elevational
and geographic clines. Rehfeldt noted,
however, that elevations above 1,800 m are
apparently of relative uniform severity, so that
additional adaptive differentiation fails to
happen. A frost in late September 1984 that
injured 2-year- old seedlings in the Coeur
d’Alene nursery of the USDA Forest Service
provided an opportunity for Rehfeldt to verify
his model. Differing injuries were inflicted on
the 5 million 2-year-old seedlings in the
nursery belonging to 179 seed lots, of which
159 were from the regions represented by the
model. Injury was recorded as a proportion of
injured trees for each seed lot, and the
resulting data then were correlated with those
predicted by the model. The statistical
significant relationship between predicted and
observed injury demonstrated the
effectiveness of the model for predicting
actual injury.

Coastal Douglas-fir
Timmis et al. (1994) built a heat-sum, fall-
hardening and spring-dehardening model to
assess the risk of frost kill to coastal
Douglas-fir in Washington and Oregon west
of the crest of the Cascades, allowing
prediction of LT50. The aim of the model
was “to provide foresters with quantitative,
operational guidelines to allocate families
within or across cur- rent breeding zones
based on frost damage risk” (Timmis et al.
1994). The model was based on the results
of freezing tests of 2-year-old seedlings over
four winters and on weather data of 80
stations on the west side of the Cascades in
Washington and Oregon from 1948 through
1990.
258 Douglas-fir: The Genus Pseudotsuga
in the western Oregon Cascades than at equivalent Bialobok and Mejnartowicz’s study included 104
elevations in the Washington Cascades. Also, an provenances from the 1966/67 IUFRO seed
apparently anomalous area for the risk of spring collec- tion covering about the same latitudinal
frost in the Coastal Ranges, between lat 45° N and range. The frost damage reported by Campbell
46° N, extends southeast into the Willamette and Sorensen was considerably more severe (with
Valley. The model indicated that the greatest risk percent damage from least to most damaged
of frost damage exists at the lower slopes of the provenances ranging from 9 to 77) than that
western Oregon Cascades. The authors concluded reported by Ching and Bever (4 to 36) and
that, “the results are in general agreement with Bialobok and Mejnartowicz (0 to 43). Considering
observations that spring frost damage is more plants at the same stage of bud devel- opment for
common than fall damage, that Oregon has more each degree of latitude separating seed sources,
higher-risk sites and that some damaging frosts the percentage of plants frosted differed by 4%
known to us are among those predicted.” in Campbell and Sorensen’s study, and by 2% in
their reanalysis of the Ching and Bever data.
Geographic Variation
Likewise, observations of frost damage in German
Geographic variation of cold hardiness in provenance tests indicated a north-south cline in
Douglas- fir has been shown in both its coastal frost resistance of coastal Douglas-fir. Frost
and interior variety. Geographic variation in frost damage to 1- and 3-year-old seedlings,
hardiness appears to be an adaptive response to representing 72 prov- enances, occurred in the
factors of the operational environment, especially three nurseries where they were raised for the
photoperiod, temperature, and moisture. Campbell 1970 international Douglas-fir provenance trial in
and Sorensen (1973) provided probably the first the Federal Republic of Germany (Kleinschmit et
experimental proof of a north-south latitudinal al. 1974). Although the extent of frost damage
gradient in frost hardi- ness of coastal Douglas-fir varied among nurseries located in the north- west,
based on natural frost damage. They scored first- southwest, and southeast of the country, plants
year seedlings growing in raised nursery beds from northern seed sources sustained less damage
(cold frames) at Corvallis, Oregon (lat 44°30’ N, than did those from southern seed sources in each
long 123°40’ W; elevation 90 m), for damage of the nurseries. Compared with British Columbia
sustained by a night frost in mid- October 1969. provenances, those from California suffered
The plants stemmed from seed col- lections in nearly 100% more damage, provenances from
five stands, each along two latitudinal transects Oregon about
from northern Washington to southern Oregon. 20%, and those from Washington about 10%.
The western transect followed the Pacific coast Damage by a late frost to 3-year-old seedlings
along the 124° meridian; the eastern transect was in a provenance trial by the Federal Institute of
about 1° (about 80 km) farther inland, skirting the Forest Genetics and Forest Tree Breeding at
east side of the Pacific coast mountain ranges. Schmalenbeck near Hamburg provided additional
Stands in which the collections had been made evidence for a north to south cline of frost
were separated by approximately one degree resistance in coast- al Douglas-fir (Reck 1978).
increments from lat 48° to 43° N. The seedlings repre- sented 85 provenances
The frost damage observed indicated distinctly whose seed sources ranged from lat 51° N in
lower frost hardiness in seedlings from the British Columbia to lat 42° N in California. The
southern than the northern seed sources. The gradient of damage with 33% at 51° N to 42% at
authors pointed out that their findings confirmed 44° N was nearly linear. From lat 44° N, the
observations on frost damage of 1-year-old percentage of damage increased abruptly to 58%
seedlings in nurseries at Corvallis (Ching and at 43° N and to 63% at 42° N.
Bever 1960) and at Kornik, Poland (Bialobok and Tallies of new shoots killed by late frosts in
Mejnartowicz 1970). The study by Ching and 1968, 1969, and 1973 in three provenance
Bever involved 14 provenances cov- ering a plantations in Michigan provided another example
latitudinal range from 49°19’N to 42°20’; of geographic variation in cold hardiness of
Douglas-fir (Steiner and Wright 1975). These
plantations were established in 1961 with 1-2
stock for testing provenances from
Chapter 12. Adverse Abiotic Factors 259

128 seed sources throughout the species’ natural reported on damage by a late October frost to
range (Wright et al. 1971). These authors divided first- year seedlings planted in cold frames at
Douglas-fir into nine groups, one coastal and Corvallis, Oregon. The seedlings stemmed from
eight interior, based on genetic similarities. seed collect- ed in six breeding units in
Steiner and Wright (1975) found provenances southwestern Oregon. Seventy-two percent of
belonging to the northern groups to be seedlings from the coastal breeding unit suffered
consistently less damaged than those from the damage; damage to seedlings from each of the
southern groups (Table 12.3). Date of flushing other units, 80 to 170 km further inland, averaged
recorded for trees at one of the plantations in 1973 less than 15%.
was shown to be highly correlated with the Larsen (1978b) investigated frost hardiness
amount of frost damage (r = −79). They of 4-year-old seedlings along two elevational
concluded that this trait is the major determinant tran- sects, one in northern and the other in
of suscep- tibility to late frost because southern Washington, on the west slope of the
provenances with the earliest bud break were the Cascades. The sample from northern
least susceptible to frost damage (Table 12.3). Washington comprised five seed sources and
Larsen and Ruetz (1980) used a transect nearly that from southern Washington four seed
along lat 44°25′ N from long 121°45′ W on the sources. Seedlings were tested for cold
Pacific Coast in the west to long 123°40′ W at the hardiness in fall, winter, and spring with
Ochoco Mountains in the east to study frost artificial freezing in the laboratory. He used
resistance of Douglas-fir along a longitudinal LT50 of needles
gradient. The 11 provenances represented by the and buds as a measure of hardiness. The results
study stemmed from a seed collection made in did not show a straight linear relation of increas-
1976 in Oregon by the Bayrische Landesanstalt ing frost hardiness with increasing elevation along
für forstliche Saat und Pflanzenzucht at either transect. Plants from the low elevation seed
Teisendorf, Bavaria. One-year-old seedlings were sources (350–500 m) were very resistant to early
subjected to freezing tests in the laboratory in and late frost. Seedlings of sources from
fall, mid-winter, and spring using elevations between 500 and 650 m were the least
LT50 of buds as a measure of frost resistance. The resistant to early, winter, and late frost. Seedlings
tests indicated that resistance to fall and winter from high elevation seed sources (800 m) were
frost increased with increasing distance from the particularly resistant to fall and winter frost.
coast. The tests, however, did not demonstrate Larsen sought to explain the relative high frost
an increasing resistance to late frost with tolerance of the low elevation provenances as an
increasing distance from the coast. adaptation to poor cold air drainage at their sites
Another indication of a west-east gradient in of origin.
resistance to fall frost of coastal Douglas-fir came Reck (1978) detailed the amount of damage to
from a study by Loopstra and Adams (1989). 3-year-old seedlings by a late frost according to
They el- evation of their seed source. The trees
belonged to 41 Washington and 23 Oregon
provenances of coastal

Table 12.3 Performance of Douglas-fir provenances within 8 groups at the Kellogg plantation in 1973 at age 12 from seed. The COAST
group is not included because most coastal Douglas-firs, unable to adapt to Michigan’s climate, died (modified from JW Wright et al.
1971).
Susceptibility to frost Time of bud burst *
Northern Rocky Mountains – NOROC Very low 8.4
Inland Empire – ILBNEMP Very low 8.1
Central Montana – CMON Low 6.2
Alberta – ALB Medium 5.1
Central Washington – CWASH Medium 4.1
Southern Colorado – SOCOL High 3.3
Northern Colorado – NOCOL Very high 3.0
Arizona and New Mexico – ARINEM High 2.9
*1 = early, 10 = late
260 Douglas-fir: The Genus Pseudotsuga
Douglas fir from the IUFRO seed collections used March had periods of
in the 1970 provenance trial of the Federal
Institute of Forest Genetics at Schmalenbeck. He
analyzed dam- age from sea level to 1,050 m by
intervals of 150 m. Reck found an elevational
pattern of damage similar to that reported by
Larson (1978): low amounts of damage at low and
high elevations and highest percentage of frost
injury at medium elevations. Percentage of
damage at all elevational intervals was, on the
average, 20% higher for Oregon than for
Washington provenances.

Aspect of site
Abele (1909) was perhaps the first to observe that
different directions of aspect of a site can result
in differing degrees of frost damage. He reported
that damage to young plantations of Douglas-fir
in Bavaria by a late frost in spring of 1908 was
most severe on sites with exposure towards the
south and southwest. In his survey of damage by
the November 1955 freeze in the Pacific
Northwest, Duffield (1956) found that damage
was generally greatest on south and southeast
slopes. A review of frost damage to coastal
Douglas-fir in the extremely cold winters of
1928/29 and 1955/56 in central Europe (Jahnel
1959) showed that trees on sites with north aspect
had suffered remarkably little damage compared
to those on south-facing sites.

Frost-induced drought
A phenomenon referred to as frost dryness
(Schönhar 1965) or winter desiccation (Sakai
1970) causes a particular kind of frost damage. It
happens when soil is frozen and trees lose more
water through transpi- ration on sunny days
than they can compensate for because of
impeded water uptake from the frozen soil.
Such damage is common in central Europe and
becomes usually apparent in early spring. The
symptoms are a brown-red discoloration of the
foliage that frequently tends to begin at the tip
of the needles (Schönhar 1965). Young trees are
often damaged so much that they die.
Nanson (1964) surveyed frost damage to
Douglas- fir in Belgium after the winter 1962/63,
the coldest in that country since keeping of
climatic records in 1833. The soil remained frozen
until the beginning of April 1963. February and
days with below-freezing air temperatures at cause the red belt. Larsen (1981) studied the
night and above-freezing air temperatures geographic variation in resistance to frost-induced
during the day. The days had strong solar drought of Douglas-fir using growth chamber
radiation and dry east winds. Nanson found experiments. He used 2-year-
that plantations with a south- facing aspect
had suffered more damage than those facing
north. Stands stocking on shallow soils had
sustained more damage than those on deep
ones. At the Groenendal tree-nursery, soil in
the open was frozen to a depth of 80 cm, but
only to 10 cm under shelter. That led him to
conclude that a combination of three factors,
depth of frozen soil, depth of root penetration,
and dry east winds, was the cause of the
observed damage. Because the combination of
these factors leads to desiccation of plants, he
felt justified in ascribing the damage to
physiological drought rather than direct frost
injury.
Damage to Douglas-fir by frost-induced drought
has been reported from southern Germany by
Schönhar (1965) and Oeschger (1973).
Seedlings were found to be especially
susceptible to winter desiccation, so much so
that in some years, losses of more than 30%
of Douglas-fir seedlings in nurseries have not
been a rare occurrence. Likewise, consid-
erable damage has been observed in
plantations of young Douglas-fir.
Suffering from physiological drought is
not unique to seedlings and saplings. Color
change of the foliage of pole-sized and mature
Douglas-fir to red-brown has been observed
in the mountains of western North America
(Scheffer and Hedgecock 1955) and the Harz
Mountains of Germany (Puchert 1954) on
south-facing slopes in early spring. This
phenomenon has been confined to a lateral
belt, generally at elevations between 1,000
and 1,500 m, and has therefore been referred
to as “red belt.” Sakai (1970) stated that “it is
reasonable to assume that the so-called ‘red
belt’ is probably caused by intensive
dehydration which arises from a
combination of exposure to sunshine and
freezing of the soil on the southern slopes.”
He theorized that a warm, dry air wall
remaining at a definite height on south-facing
slopes for several hours to a few days in late
winter, known as a subsidence inversion, may
be the deci- sive factor acting as a trigger to
Chapter 12. Adverse Abiotic Factors 261
transpi- ration rates.
old seedlings of 40 provenances from the IUFRO
seed collection to investigate two components:
drought avoidance and drought tolerance, of frost-
induced drought. He measured avoidance of
desiccation as the rate of negative increase in bars
of water poten- tial per day, and drought tolerance
as bar of water potential resulting in 50% of plant
mortality. His findings indicated that the 40
provenances could be divided into three groups
based on their degree of drought avoidance and
drought tolerance:
1. Those with a very low degree of drought
avoid- ance and very low drought tolerance.
They were from the coastal regions and the
Cascade Ranges of British Columbia,
Washington and Oregon.
2. Those with a very high degree of drought
avoid- ance and drought tolerance. They came
from Colorado, Arizona, and New Mexico.
3. Those that were intermediate between group 1
and 2 regarding degree of drought avoidance
and drought resistance. They were from
interior British Columbia and Idaho.
Among the relatively drought-sensitive
provenances of group 1, Larsen found an increase
in resistance to frost-induced drought with
increasing elevation of origin.
Rossa and Larsen (1980) investigated in a com-
panion study the effect of cuticular transpiration
on resistance to frost-induced drought in 35 of the
40 Douglas-fir provenances used in Larsen’s 1981
study of geographic variation in resistance to
frost-induced drought. They measured cuticular
transpiration of detached twigs of 3-year-old
seedlings in a growth chamber. Based on their
cuticular drying rate, the 35 provenances could be
divided into the following groups:
1. Provenances from coastal British Columbia
including Vancouver Island, Washington and
Oregon west of the crest of the Cascade
Range, and northern California having high
cuticular transpiration rates. A definite
influence of el- evation or distance from the
coast of their seed source on cuticular
transpiration was not appar- ent within this
group of provenances.
2. Provenances from Idaho, Colorado, Arizona,
and New Mexico having very low cuticular
have allowed them to survive a normal winter
3. Provenances from interior British in their native habitat. That finding indicated
Columbia with cuticular transpiration that low N may impede cold hardening only
rates intermediate between those of with a relative overabundance of P and K. The
provenances in group 1 and 2. results of the study suggest that normal
In addition, Rossa and Larsen (1980) studied development
cu- ticular thickness and stomatal depth in
relation to cuticular transpiration in seedlings
of 10 of the 35 provenances in their study.
They found the thickest cuticula in the
provenances that were from the south- ern part
of the range of interior Douglas-fir. These
were the provenances that had the highest
resistance to cuticular transpiration.
Provenances belonging to the coastal variety
of Douglas-fir that were the most susceptible
to desiccation had the thinnest cuticula.
Provenances from interior British Columbia
had cuticula of medium thickness. Stomatal
depth was found to be greater in the
provenances of interior than coastal Douglas-
fir. The findings demonstrated that rates of
cuticular transpiration of Douglas-fir are
closely correlated with the anatomical charac-
teristics of its needles.

Nutrients
Nutrient levels are among the many factors
that influence cold hardiness of Douglas-fir.
Alden (1971) found that trees from a
plantation deficient in potas- sium and
nitrogen developed significantly lower
hardiness in the acclimation stage than did
trees from a plantation not deficient in these
nutrients. Larsen (1976) investigated the
effects of different levels of fertilization with
nitrogen, phosphorus, potas- sium, and boron
on the cold hardiness of 4-year-old Douglas-
fir seedlings. He found that an extremely low
(0.9%) and a very high (2.2%) level of N had
a detrimental effect on resistance to fall and
winter frost. He could not show any effect of
K, P, and B on cold hardiness. Larsen (1978)
hypothesized that the better survival in winter
of Douglas-fir with sufficient levels of K in
their tissues is not a consequence of higher
resistance to frost but to frost dryness.
In an experiment by Timmis (1974) with first-year
coastal Douglas-fir seedlings, those
deprived of N but receiving P and K were
unable to harden off to an extent that would
262 Douglas-fir: The Genus Pseudotsuga
of cold hardiness is probably more closely related var.
to a balance between nutrients than to the level of
any single nutrient element.
Intraspecific hybridization
Despite a long history of differentiation that may
ex- tend as far back as the Miocene, the coastal
and inland varieties readily cross, both under
controlled pollina- tion and in the wild
(Critchfield 1984). Ecklundh’s experiments
(1943) in Sweden (Schönbach 1958, p.
366) were the first to explore the possibility of
com- bining the growth potential of coastal
Douglas-fir with the frost hardiness of interior
Douglas-fir. These crosses, however, were not
successful because they only yielded empty seeds.
Nearly 20 years later, Schönbach of the
Institute for Forest Sciences at Eberswalde,
Germany, made two sets of reciprocal crossings
of coastal and inte- rior Douglas-fir. As a result,
22 combinations were available for tests in 1960
(Schönbach and Bellmann 1967), and another 35
combinations in 1964 (Braun and Schmiedel
1985). The crossing partners of the 1960 set were
four 80-year-old coastal Douglas-firs, two from
compartment 50, two from compartment 29 of the
Tharandt Forest, and two interior Douglas-firs
from a stand in the Elbsandstein region of Saxony.
The crossing partners of the 1964 set, five coastal
Douglas-firs, were from the same stands as those
of the 1960 set (Braun and Schmiedel 1985).
Trees from compartments 50 and 29 had been
chosen as crossing partners because an earlier
study of heritablility of frost hardiness
(Schönbach and Bellmann 1964) had shown
progeny from trees in compartment 50 to be more
frost hardy than progeny from trees in com-
partment 29. Hybrids of the 1960 set were planted
as 1−1 seedlings in spring 1963 at the Plaue forest
district in the Ore Mountains of eastern Germany
(Schönbach and Bellmann 1967). Hybrids of the
1964 set were planted as 1−1 seedlings in spring
1967 in the Tharandt forest district near Dresden
(Braun and Schmiedel 1985).
Tallies of frost damage and height growth mea-
surements of 1-year-old and 2-year-old seedlings
in the nursery, and of 6-year-old saplings in the
Plaue plantation, showed that the var. menziesii x
var. glauca (h1) and var. glauca x var. menziesii
(h2) hybrids but had better height growth than the
menziesii x va. Menziesii (m) hybrids. N, and from long 98°07’ to 125°40’ W. That also
Survival of g, h1, and h2 hybrids ranged from solved the problem of having both seed and pollen
96% to 100% 4 years after outplanting, parents at the right time of the year for crossing.
compared to 60% for the m hybrids
(Schönbach and Bellmann 1967). Survival of
trees in the Plaue plantation was 90% for the g
hybrids and 65% for the h1 and h2 hybrids,
but only 10% for the m hybrids at age 18 from
seed. Measurements 6 years later showed the
best trees of some h1 and h2 hybrid progenies
had attained a height of 21m and a dbh of 24
cm (Braun 1988). Mean tree height at the
Tharandt plantation at age 18 from seed
ranged from 11.45 m for the hybrids from one
of the h1 combinations to
5.45 m for hybrids from one of the m
combinations. Twelve of the progenies, all of
which were h1 and h2 hybrids, ranked above
the plantation mean of 9 m. Both the h1 and
h2 hybrids showed—in general— better
growth than comparable m and g hybrids. The
m hybrids whose parents were relatively frost
hardy (Schönbach and Bellmann 1967) showed
better height growth than m hybrids with less
frost hardy parents. The g hybrids were the
slowest-growing hybrids (Braun and
Schmiedel 1985). The best hy- brids in the
Tharandt plantation had produced, at age 16, a
volume of 140 m3/ha. That was twice the
volume for site class I in the Germany yield
tables of Hengst and of Bergel (Braun 1988).
Referring to the performance of these hybrid
progenies, Braun suggested (1988) a rotation
of 70 years for stands established with hybrid
seedlings, assuming such a rotation length
would yield 1,400m3/ha. But he also pointed
out the main problem of using hybrids in
reforestation practice, namely the difficulty of
mass producing hybrid seedlings.
Duffield (1950) was probably the first in
North America to suggest hybridization as a
means for im- provement of Douglas-fir. Orr-
Ewing (1973) started a Douglas-fir arboretum
in 1958 on Vancouver Island, British
Columbia to provide as wide a gene pool as
possible for intraspecific crosses. In 1973, the
arbo- retum already contained 216
provenances and 121 clones collected
throughout the natural range of the species.
Their sites of origin ranged from 30 m to
3,300 m elevation, from lat 19°40’ to 55°05’
Chapter 12. Adverse Abiotic Factors 263
their
The first 44 intraspecific crosses made by Orr-
Ewing (1966b) were limited to pollen parents
from coastal Douglas-fir ranging from northern
British Columbia to northern California. The
subsequent 114 crosses included pollen parents
from both the coastal and interior variety. Orr-
Ewing et al. (1972) established 28 test sites on
Vancouver Island and the lower mainland so that
every cross could be planted under a wide range
of climate conditions for the assessment of
genotype/environment inter- actions. Orr-Ewing
et al. (1972) concluded that the initial result
indicated the absence of an incompat- ibility
barrier preventing successful crosses between
Douglas-fir separated by thousands of kilometers
and growing in completely different
environments. That was clearly demonstrated by
the cross be- tween a maternal parent from
Madera, Chihuahua (29°10’N) and a paternal
parent from Fort St. James, British Columbia
(54°30’N). Although the distance between their
geographic origins is nearly 26 degrees of
latitude, viable seed was nevertheless obtained.
Intraspecific crosses, however, have definite
limits for obtaining increased growth (Orr-Ewing
et al. 1972). Crosses with some of the Washington
and Oregon paternal parents were very promising,
but those with California paternal parents were
not very successful. Crosses with paternal parents
from the interior range of Douglas-fir gave no
positive growth results in the maritime climate of
south- western British Columbia. Such hybrids,
however may be of value in the interior of British
Columbia (Orr-Ewing et al. 1972).
Rehfeldt (1977) began a study in 1971 to ex-
plore the potential of intervarietal hybridization
for improving Douglas-fir in the northern Rocky
Mountains. He produced 70 hybrid families by
pollinating 20 interior Douglas-firs from 2 Idaho
provenances with pollen from 25 coastal Douglas-
firs representing 3 Oregon and 1 British Columbia
provenance. Each parental tree was also
represented in the study by seedlings derived from
wind pollina- tion in their native stands. The
result was 33 half-sib families representing
parental lines.
Seeds were sown in October 1971 at the Priest
River Experimental Forest nursery and grown for
3 years. Performance of the hybrid families and
either the parental provenance of the specific
parental lines showed that the growth parental tree. In addition, variance within families
potential of hybrids was generally superior to was high.” He saw the solution for an expedient
that of the interior variety, but was similar to and safe means of using hybridization in breeding
that of the coastal variety. The harsh winter of of interior Douglas-fir in
1972, which brought tempera- tures as low as
−26°C, led to high rates of mortality. The rate
of survival was 44% for seedlings of hybrid
and interior origin, but only 9% for those of
coastal origin. Little additional mortality
occurred during the next 3 years, suggesting
that the capability of hybrid families to
survive under the severe climate of the
northern Rocky Mountains approaches that of
the interior variety (Rehfeldt 1977). Surviving
seedlings were planted in 1975 as 2−1 stock
in row plots on a site at 1,036 m elevation
near Grangeville (lat 46° N, long 116° W) in
northern Idaho.
Caused by the losses of seedlings at the
nursery, the Grangeville plantation of 3,025
trees contained an unequal number of crosses
per parent, an unequal number of plots per
cross, and an unequal number of seedlings per
plot. Because these imbalances precluded
statistical analyses, Rehfeldt compared
survival and height of trees 10 years after
planting without regard for statistical
probabilities. Survival averaged 58% for
hybrids, 63% for open-pollinated interior
parent lines, but only 20% for open-pollinat-
ed coastal parental lines. Average height for
hybrids was 222 cm, 128 cm for interior
parental lines, and 104 cm for coastal parental
lines. The performance of individual hybrid
families varied considerably. Survival ranged
from 34% to 80% and mean height from 140
cm to 335 cm. Thus after 10 years in the field,
hybrids essentially equaled the survival of
inte- rior parental lines, but were almost twice
as tall. The results of Rehfeldt’s study
demonstrated the high potential for
intervarietal hybridization to increase the
productivity of Douglas-fir on sites with a
harsh climate, as had the earlier German
studies initiated by Schönbach in 1960.
Rehfeldt (1986b), however, pointed out a
problem faced by using hybridization for
improving productivity of interior Douglas-
fir. He found that “the performance of a
hybrid fam- ily could not be predicted from
264 Douglas-fir: The Genus Pseudotsuga
the selection of superior hybrid trees without heaving and will die usually after such exposure
regard to parentage, and then backcrossing these (Hermann 1990).
selected trees to a large number of trees of the
interior variety.

Interspecific hybridization
Duffield (1950) attempted unsuccessfully in
1947 to cross Douglas-fir with bigcone
Douglas-fir as pollen parent. The 3 apparently
sound seeds, out of a total of 438 seeds from the
cross, failed to germinate. Ching (1959) made
the first successful cross in 1956. The female
parents were four trees in a 30-year-old stand in
the Oregon Coast Range. The pollen parents
were bigcone Douglas-firs on Baldy Mountain in
Los Angeles County, California. Ching
concluded that the crossability of the two
species is rather low, although the cross yielded
some fertile seeds. Orr-Ewing (1966b) made
some crosses in 1962 on seven Douglas-firs
growing on Vancouver Island, and again on
seven other trees in 1964 with bigcone Douglas-
fir as pollen parents. The 1962 cross yielded
only empty seeds, and of the 14,199 seeds
extracted from cones resulting from the 1964
crosses, only 8 seeds were viable. Orr-Ewing
considered the difference in chromo- some
number—bigcone Douglas-fir has 2n = 24
chromosomes, Douglas-fir has 2n = 26 chromo-
somes—as the probable explanation for the
evident incompatibility in this cross.
Pseudotsuga wilsoniana, the Formosan Douglas-
fir, is the only known Asiatic species to have been
crossed with Douglas-fir. According to Orr-Ewing
(1966b), Roy Silen, a USDA Forest Service
geneticist, attempted the first—but unsuccessful—
cross with Formosan Douglas-fir as the pollen
parent in 1962. An attempt in 1963 by Orr-Ewing
(1966b) to cross the two species did not succeed,
yielding only empty seeds. His conclusion, that
the prospect of interspe- cific crossing with
Douglas-fir is not particularly promising, has so
far not been disproved.

Frost heaving
Repeated freezing and thawing of the soil causes
frost heaving. The expansion and contraction of
the soil slowly pulls plants out of the ground,
leaving roots partially or completely exposed.
First-year seedlings are most susceptible to frost
Hermann has also noted 2−1 Douglas-fir development at elevations between 260 m and
seedlings lifted out of the ground by frost 550 m. All trees planted with a spacing of 1.2 m2
heaving in a newly established plantation, and
however. 1.5 m2 suffered greatly from snow breakage in the
winter 1909/10; those planted at a spacing of 3.0
Snow, Ice, and Hail m2 and 5.0 m2 escaped damage. Losses of up to
Snow, ice, and hail storms are climatic events 40% of
that may cause serious damage to Douglas-fir.
Heavy snowfalls in the late winter of 1963/64
and in early spring of 1964 on upper slopes of
the Oregon Cascade Range inflicted much
damage to young, mixed- conifer stands
(Williams 1966). Leaning, bent, and fractured
stems and broken branches were common for
trees 1.2 and 6.0 m in height, but few trees
below
1.2 m high suffered damage. Trees of saw
timber size were not injured, except for a few
broken branches. Douglas-fir suffered more
damage than any of its associated species. The
susceptibility of Douglas-fir to snow damage
at high-elevation sites makes the
establishment of Douglas-fir monocultures a
ques- tionable practice in cutover areas of
upper slope forests in the Oregon Cascade
Range.
Data collected from 32 permanent sample
plots in forests of the Oregon Coast Range
showed that young stands of Douglas-fir
suffered severe snow damage at elevations
above 300 m in the winters of 1964/65 and
1965/66, and above 760 m in the winter
1968/69 (Kangur 1973). His study
demonstrated that trees that were widely
spaced at seedling and sapling stages
sustained less damage than did trees in
densely stocked stands. The results also
indicated that the degree and time of thinning
can influence the amount of damage. Twenty-
one-year-old stands, thinned 2 years before the
snowfall suffered severe damage, but damage
was very light in adjacent stands thinned 6
years before the snowfall. Both Williams
(1966) and Kangur (1973) recommended
early thinnings to make stands more resistant
to snow breakage.
Puchert’s (1954) account of snow damage in 21
stands of coastal Douglas-fir (established
from 1880 to 1890 in the Harz Mountains of
central Germany) described the effects of
snow damage on long-term stand
Chapter 12. Adverse Abiotic Factors 265

trees in 20- to 40-year-old stands did not


adversely affect stand development. Snow break Stand C, covering 2.025 ha, had been formed by
had the same effect as a heavy thinning. Puchert the slow encroachment of Douglas-fir beneath an
mentioned a stand at 360 m elevation, planted at oak overstory. As the oak gradually died out,
1.0 × 1.5 m spacing, that suffered thrice from Douglas-fir filled the holes and created an
much snow breakage until age 40. This stand had, unevenaged stand. An improvement cutting in
at age 69, a volume of 551 m3, which Puchert half of the stand in 1935 removed 176 trees. An
considered an indication that Douglas-fir can unusually heavy fall of wet snow in 1936 broke
endure a considerable amount of snow damage in off 44 trees in the freshly cut half. Fifty-five of
its youth without an adverse effect on volume the remaining 73 trees in that half of the stand
production later in its life. He concluded that were broken or uprooted by the 1942 break in
Douglas-fir is most susceptible to snow break, 1936, and 63 trees of the residual stand had ice
particularly in densely stocked stands, between damage in 1942. The volume lost by ice and snow
ages 20 and 40. damage in the three stands amounted to about
Freezing rain forms layers of ice on trees that one- fourth of the original volume (Table 12.4).
bends and breaks stems and branches. Three pub- Freezing rain in January 1970 on both sides of the
lished records (McCulloch 1943, Anonymous Columbia Gorge led to losses that ranged to 80%
1971, Russell 1971) documented the extent of of trees in up to 40-year-old Douglas-fir stands on
the Oregon side of the Gorge (Anonymous 1970).
known damage from ice storms in Douglas-fir
Damage was estimated to have occurred on
forest of the Pacific Northwest. McCulloch (1943)
16,200 ha. Other factors sometimes contributed to
invento- ried the damage from the January 1942
the ob- served damage. A damage survey in a
ice storm in three second-growth stands in
Douglas-fir stand in the Cascade Range of
western Oregon. Stand A, age about 90, average
southern Washington after the January 1970 ice
height 33 m, average dbh 34 cm, covered 4.05 ha.
storm revealed large num- bers of uprooted trees
Half of the stand had remained uncut; the other
whose roots had already been weakened by
half had been thinned twice, in 1933 and 1938.
infection with Phellinus weirii before the ice
The two cuttings removed 371 mostly suppressed
storm (Russell 1971).
and intermediate trees and thus did not create
Damage caused by hail is rare, but can be
large holes in the canopy. Yet this partially cut
disas- trous. A hail storm with hail stones of
stand lost 267 trees by breakage and uprooting in
unusual size near Angers, France, in August 1944
the ice storm compared to a loss of 124 trees in
mutilated trees in young Douglas-fir plantations to
the uncut half. Stand B, age about 80, average
the extent that they had to be replanted (Vazeilles
height 27 m, average dbh 26 cm, covered 3.24 ha.
1946).
A 1941 thinning in half of stand B had removed
131 dominant and codominant trees, which Drought
resulted in a sizeable opening of the canopy. The
Summer drought is common in many parts of the
thinned part lost 471 trees through breakage and
Douglas-fir region and has been responsible for
uprooting by the ice storm, but the uncut part lost
some failures of natural and artificial regeneration
only 101 trees.
of Douglas-fir. Outright kill of older trees,
however,

Table 12.4 Volume lost in three Oregon Douglas-fir stands to ice damage in January 1942. Data from McCulloch (1943), converted to metric
measurements.

Stand A Stand B Stand C


m3 % m3 % m3 %
Original volume 994 100.0 711 100.0 405 100.0
Volume cut 169 16.6 168 23.6 258 63.7
Lost in storm 236 23.7 145 23.1 117* 28.8
Residual volume 589 59.7 398 53.3 30 7.5

* Snow and ice storms.


266 Douglas-fir: The Genus Pseudotsuga
appears to be rare. Childs (1960) reported on the moist and dry sites in Oregon and Washington. The
effects of the unusually hot summer of 1958 and 1968 investigation used 2- and 3-month-old seedlings;
the unusually dry summer of 1959 on Douglas-fir. the 1970 study
Death of young trees was common in a few
localities in northwestern Washington at the end
of the 1958 growing season. He observed
drought damage in both years in Washington and
Oregon manifested by partially dead crowns.
Top kill was confined, however, almost entirely
to Douglas-fir saplings and poles. Drought
damage was especially conspicuous on clay soils
along the east side of the Mt. Hood National
Forest, where such damage had also oc- curred
during the drought of the late 1920s and early
1930s. Childs (1960) stated that “on fair to good
sites mortality in young stands is rarely extensive
enough to impair stocking. Of the economic loss
attributed to drought, by far the greater part has
undoubtedly resulted from inconspicuous but
general decrease in current increment throughout
most of the region.” Child’s observations of
drought damage to Douglas-fir in the Pacific
Northwest were followed by several greenhouse
and laboratory studies of dif- ferences in drought
resistance and drought avoidance between
Douglas-fir of different geographic origins.
Greenhouse studies by Ferrell and Woodard
(1966) and Pharis and Ferrell (1966) showed
seedlings from inland sources to be more
drought resistant than those from Pacific Coast
sources. They considered the higher survival of
interior Douglas-fir in their tests to reflect true
drought hardiness rather than drought avoidance.
Their findings indicated that Douglas-fir from
xeric habitats is more drought hardy than is
Douglas-fir from mesic habitats. Zavitkovski and
Ferrell (1970) theorized that “different kinds of
natural selection operate in these 2 environments:
drought in xeric habitats will favor drought-
resistant individuals, but in mesic environments
drought resistance is secondary to other
ecological factors in natural selection. This
circumstance suggests that seedlings from these 2
environments may differ in
their physiological responses to drought.”
That assumption was confirmed in studies by
Zavitkovski and Ferrell (1968, 1970) on the
effects of drought on photosynthesis,
transpiration, and respiration of seedlings from
used 2-year-old seedlings. Although burst showed that the dry-source progeny had
photosynthet- ic, respirational and much earlier and more complete bud burst than
transpirational rates showed similar declines the wet-source progeny. Measurements of root
in all seedlings with decreasing soil moisture, systems of live and dead seedlings excavated
mesic source seedlings had significantly
higher transpirational rates than did xeric-
source plants. But mesic-source seedlings had
considerably higher photosynthetic rates at
soil moisture tensions between 1 and 15 atm.
Zavitkovski and Ferrell (1970) considered that
as a range of soil moisture stress ex- isting
probably under natural conditions most of the
time during the growing season in the
Douglas-fir region. They suggested that the
high transpirational rates of mesic-source
seedlings that resulted in more rapid
exhaustion of available soil moisture would
be detrimental to their establishment only on
extremely dry sites. On moderately droughty
sites, however, their ability to maintain high
photosynthetic rates under low soil moisture
stress may overcome the disadvantage of high
transpirational rates.
Heiner and Lavender (1972) used a different
approach to investigate the response of mesic-
and xeric-provenance Douglas-fir to drought-
caused stress. They collected seed from five
trees on a site with annual rainfall of 508 mm
in southern Oregon, and from many trees on a
site with annual precipi- tation of 1,542 mm in
the northern Oregon Coast range. Seedlings
were raised in a nursery in the Willamette
Valley. Seedlings from dry and wet seed
sources were transplanted at age 2 into a
lysimeter. The soil was wetted to field
capacity at the time of transplanting and then
allowed to dry naturally during the growing
season. Depletion of soil mois- ture at depths
of 10, 30, and 60 cm was measured
fortnightly. A plastic canopy suspended above
the lysimeter prevented rains from moistening
the soil. Additionally, seedlings from each
seed source were maintained in the nursery
and kept well watered during the growing
season. None of these seedlings died in their
3rd growing season. In the lysimeter, survival
of progeny from each of the five trees from
the dry site in southern Oregon ranged from
55% to 70%, compared with 16% for progeny
from the wet site in northwestern Oregon.
Weekly tallies of termi- nal and lateral but
Chapter 12. Adverse Abiotic Factors 267
and
at the end of the growing season showed that the potassium supply on drought hardiness of 2-year-old
progeny of the five trees from the dry site in Douglas-fir from Snoqualmie, Washington, grown
south- ern Oregon had mean root lengths ranging
from 35 to 42 cm, compared with 31 cm for the
progeny of trees from the wet site in northwestern
Oregon. None of the dead trees from all seed
source had roots longer than 23 cm. Seedling
survival apparently was largely dependent upon
the ability of roots to penetrate below a depth of
30 cm. The correlation found between early bud
break and survival in turn reflected a correlation
between early growth and survival. The principal
survival mechanism identi- fied by Heiner and
Lavender (1972), vigorous early root growth, is
actually drought avoidance rather than a drought-
resistance mechanism.
White (1987) studied differences in drought
tolerance between populations of Douglas-fir in
southwestern Oregon, where summer drought in
particular limits the success of natural and
artificial regeneration on some sites. For his
study, White used seed from 72 open-pollinated
families from 2 parent trees each at 36 locations
throughout southwestern Oregon. Sample
locations were 61 km to 162 km from the Pacific
Ocean and were between lat 42°00’ and 43°12’ N,
and at 475 m to 1,630 m altitude. A drought
regime was imposed in a growing room, a
greenhouse, and an outdoor cold frame on some
of the seedlings in their second growing season,
and on others in their third. Seedling survival was
measured under an extended soil drought
designed to simulate that which often occurs in
southwestern Oregon: adequate moisture early in
spring, but little or no rain after budburst. Thus,
watering was discontin- ued after budburst. The
results were similar in the three test environments.
They showed that drought- tolerant populations
were from higher elevations, and to a lesser
extent, from drier sites. The popula- tions from
higher elevations had an earlier budburst than
those from lower elevations. Early budset was
strongly correlated with increased drought
tolerance in White’s study. He considered the
early budset, and hence early entry into
dormancy, to be a possible explanation for the
increased drought tolerance of populations from
higher elevations.
Larsen (1983) studied the effect of nitrogen
loess soil of the site. Evapotranspiration during
under 11 different levels in a growth room. He the 1975 growing season (210 days) was on
found that K had a large positive effect on average 2.38 mm/day. Evapotranspiration in the
drought hardi- ness. By contrast, an increasing growing season (202 days) of the much drier year,
supply of N caused a significant decrease in 1976, was only 1.46 mm/day. Moisture stress
drought tolerance. experienced by the tree during the
Development of the pressure bomb
(Waring and Cleary 1967) made possible the
measurement of plant moisture stress (PMS)
in Douglas-fir and thus to as- sess directly the
level of moisture stress, which was
particularly helpful when working with
seedlings and saplings. Waring and Cleary
(1967) showed that with adequate soil
moisture, PMS in Douglas-fir var- ies in the
course of day, with its peak shortly before
10:00 a.m. and approaching a minimum by
8:00
p.m. They also demonstrated that stresses in
plants are considerably higher than in the soils
they grow on by showing that on a bright day
trees growing on soils near field capacity can
have stresses of 20 bars. The soil moisture at
field capacity is 0.3 bar. The nature of soil on
which Douglas-fir grows also has a bearing on
the severity of PMS. Cleary (1969) followed
PMS in Douglas-fir on different sites as the
dry season in Oregon progressed. As early as
the end of June, he found little night-time
recovery of PMS on a very coarse granitic
soil. By contrast, in a fine-textured soil, there
was considerable recovery at night as late as
August, although dawn minima in July and
August exceeded 30 bars. The height of
Douglas-fir also influences PMS. Waring and
Cleary (1967) pointed out that when 1-m-tall
trees had stresses nearing 40 bars, 25-m-tall
trees had stresses of 20 bars, presumably
because the larger trees could tap the deeper
soil layers where more water was still
available.
A study by Borer (1982) on the effect of drought
on a mature Douglas-fir in a stand in
Switzerland showed that drought may not
result in visual dam- age, but nevertheless
may cause damage through the reduction of
growth. He investigated the water uptake
and evapotranspiration of an 85-year-old
Douglas-fir in 1975 and 1976. The tree was
43 m tall and had a dbh of 81 cm. Its root
system ex- tended to a depth of 2 m in the
268 Douglas-fir: The Genus Pseudotsuga
drought period from June to mid-July did not permits establish- ing precisely the years of fire injury
result in visible damage. Borer attributed the or tree origin.
lack of vis- ible damage to a water reserve large
enough in the 2-m layer of soil occupied by the
tree’s root system to enable the tree to come
through the drought without lasting damage. He
stated, however, that biomass production was
reduced during the period of drought. Borer
estimated wood production of the Douglas-fir
tree to be about 0.3 m3 in 1975, based on its
consumption of water. For the production of
that amount of wood, the tree had to take up
90,000 L of water during the growing season.
More than 99% of this amount of water was lost
by transpiration.

Fire
Wildfire has been a major natural disturbance in
the Douglas-fir forests of western North America.
The presence of charcoal layers below ash from
the eruption of Mount Manzama 6,700 years BP
indicate that wildfire has been a primary
disturbance mecha- nism in the Cascade Range
for at least 10,000 years (Morrison and Swanson
1990). Extensive fire activity occurred at least
every decade or two in the Inland Northwest
(eastern Washington and Oregon, western
Montana and Idaho) between the 1500s and the
early 1900s (Barrett et al. 1997). The annual area
burned during these 350 years has been estimated
at about 146,800 hectares of the nearly 7.7 million
hectares covered by interior Douglas-fir and
western larch. Three different fire regimes, based
on their se- verity, are recognized for describing
fire events (Bradley et al. 1992). Low severity
fires are surface fires that burn litter, duff, loose
woody debris on the forest floor, and
undergrowth vegetation. High severity fires cause
high or complete mortality in an overstory stand
of trees, and are often referred to as “stand-
replacing” fires. Moderate severity fires define a
broad range between those two extremes.
Methods used for the reconstruction of fire his-
tory include fire scars found on cat-faced trees
and on cross sections of stumps, cores from live
trees, post-fire regeneration age classes, and
written re- cords (Morrison and Swanson 1990).
Cross-dating, the matching of tree-ring patterns to
determine ab- solute dates for tree-ring series,
These can then be used to date wildfires combinations of frequent low- and moderate-
(Weisberg and Swanson 2001). severity fires and infrequent stand-replacing fires.
Views have changed since the end of the These studies also suggest a general pattern of
20th century on the frequency of fire increased frequency
regimes of differ- ent intensity in the
montane forests dominated by coastal
Douglas-fir in the area west of the crest of
the Cascade Range and from northern
California to Washington. According to
Wetzel and Fonda (2000), for over half a
century, Douglas-fir forests on the Pacific
slope were depicted as supporting infrequent
stand-replacing fires. Mean fire return
intervals (FRI) of 200–400 years were
commonly estimated for Douglas-fir forests in
the Pacific Northwest, largely based on dates
of stand establishment after high severity
fires. (FRI is the mean time span between
fires, specific for a given unit of land,
vegetation type, or region).
But the fire regime associated with forests
dom- inated by Douglas-fir is far more
complex than originally reported. Studies of
the fire history in Douglas-fir forests of the
Pacific Northwest con- ducted since the late
1980s have shown that fire fre- quency is
much shorter, and fire severity much less,
than previously thought. Yamagushi’s (1986)
study of the fire history in old-growth
Douglas-fir stands northeast of Mount St.
Helens revealed a record of large stand-
replacing fires and relatively frequent low
severity fires following the eruption of the
vol- cano in 1480. The frequency of these low
severity fires was one fire per 40 to 50 years
during the first 150 years of stand
development and one fire per 125 to 150 years
thereafter. That frequent moderate and low
intensity fires were part of the fire regime in
the Douglas-fir forests of the Cascade Range
became even more apparently from the
reconstruction of the fire history in two 1,940-
ha areas in the central- western Cascade
Range of Oregon by Morrison and Swanson
(1990). From their study and several other
studies, it has become clear that Douglas-fir
forests from the northern Cascades in
Washington (Agee et al. 1990) through the
central Cascades (Wallin et al. 1996, Cissel et
al. 1998), the Klamath Mountains (Taylor and
Skinner 1998), and northern California
(Brown et al. 1999) have experienced
Chapter 12. Adverse Abiotic Factors 269

and decreased severity of natural fires from north into series and habitat types in the northern Rocky
to south. Mountains. The Douglas-fir series in the forests
The montane forests of the northeastern west of the Continental Divide forms an
Olympic Mountains are dominated by Douglas-fir especially broad zone in Montana and central
stands es- tablished after past burns (Fonda and Idaho. The four principal tree species of the series
Bliss 1969). A 600-year fire history of Douglas-fir are ponderosa pine, Douglas-fir, western larch,
forests developed by Wetzel and Fonda (2000) for and lodgepole pine (Pinus contorta). Surface fires
a 2,500 ha drainage in the northeastern Olympics of low to moder- ate severity were very common
revealed that periods with many small-scale, and in the Douglas-fir series. Suppression of fire after
low- and moderate 1900 resulted in an accumulation of surface fuels
-severity fires were interrupted by two periods of and the develop- ment of a Douglas-fir understory
stand-replacing fires in 1687–1720 and 1897– on dry and moist sites (Arno 1988). In many dry
1904. That fire history shows that small patchy Douglas-fir habitats, where ponderosa pine is
fires were much more common in the eastern seral, the establishment of an understory of
Olympics than previously thought. The Douglas-fir would lead in time to replacing the
reconstruction of the fire record in the eastern shade-intolerant ponderosa pine with a dense,
Olympics indicated a low in- cidence of fires disease- and insect-prone Douglas-fir climax
during the little Ice Age, a climatic period of low (Arno et al. 1995). Consequently, stand-
temperatures extending through the 17th and 18th replacement fires became common (Harrington
centuries to the mid-19th century (Henderson and 1991, Bradley et al. 1992). Stand-replacement
Brubaker 1986). The marked in- crease in the fires are sometimes followed within a few years
1850/59 decade appears to signal the return of a by reburns. Perhaps the best known example is
drier and warmer climate at the close of the Little the 1953 Tillamook burn in Oregon that destroyed
Ice Age (Wetzel and Fonda 2000). about 104,000 ha of old- growth Douglas-
Interior Douglas-fir in the forests east of the fir/western hemlock forest. The area experienced a
crest of the Cascades and in the Rocky Mountains 81,000 ha reburn in 1939, and a second reburn of
grows as a climax of major seral species on a nearly 73,000 ha in 1945 (Anonymous 1966).
wide variety of sites, with climates ranging from Reburn fires can be more intense because recently
warm semiarid to inland moist to high-elevation burned stands often have greater amounts of fuel
cold (Pfister et al. 1977). In the eastern mountains than mature forests (Agee and Huff 1987).
of Washington and Oregon, frequent low-severity The reestablishment of Douglas-fir after initial
fires for centuries maintained open forests with burns or reburns was often delayed by a limited
large, widely spaced, predominantly fire-tolerant seed source, invasion of brush, altered soil
trees: that is ponderosa pine (Pinus ponderosa), nutrient status, and damage caused by wildlife.
Douglas-fir, and western larch (Larix occidentalis) Franklin and Hemstrom (1981), in a discussion of
(Everett et al. 2000). Changes in forest the impact of fire on succession in the coniferous
management practices, notably fire suppres- sion, forests of the Pacific Northwest, hypothesized that
led to a marked decline in fire events since about repeated wild fires were responsible for the slow
1900. That enabled seedlings and saplings, establishment of many of the old-growth forests
particularly of fire-sensitive species such as grand that originated about 500 years ago in the Cascade
fir (Abies grandis) to invade and persist beneath Range. Gray and Franklin (1967) studied the
the overstory of the open forests; whereas the effects of multiple fires on the reestablishment of
invaders would have previously been eliminated Douglas-fir in a 16,000 ha watershed in the
by frequent low-severity surface fires. The result Cascade Range of southwestern Washington.
was the de- velopment of ladder fuels and large Most of the Douglas-fir in the water- shed was
fuel buildups, leading to moderate- and high- probably older when a fire in September 1902
severity fires in forests that historically did not killed most of the trees in the drainage. The 1902
experience them (Agee 1994, Everett et al. 2000). fire was followed by reburns in 1919, 1927, and
Pfister et al. (1977) distinguished forest zones 1932. Although of catastrophic intensity, they
defined by the potential climax species arranged were much smaller than the initial 1902 burn.
Their find-
270 Douglas-fir: The Genus Pseudotsuga
ings indicate that degree of intensity of burn and initiated by Norum (1975) and observations 8
the site characteristics are among the factors years after the 1973 spring and fall burns for the
determin- ing the speed of regeneration. construc- tion of their model. Of the 166 trees in
Reestablishment of Douglas-fir on single-burn the sample, 83 were dead after 8 years. The
sites took less time than on reburn sites. majority (70 trees) died in 1974 and 1975.
Reestablishment of Douglas-fir on single-burn Another 13 trees died in the fol- lowing 6 years.
sites took about 12 years on its wetter sites Their analysis showed that survival decreased
compared to 24 years or more on its drier sites. with increasing scorch height, percent of crown
The damage fire causes to Douglas-fir depends scorch, and number of quadrandts with dead
mainly on the intensity of fires and the tree’s age. cambium at 1.4 m bole height, but increased with
Morrison and Swanson (1990) attributed the larger diameters. Ryan et al. (1988) found that
ability of old-growth stands of Douglas-fir to percentage of crown scorched was a better predic-
withstand re- peated low- to moderate-intensity tor of tree mortality than scorch height,
fires to the stand’s physical characteristics, confirming findings by other workers (Peterson
namely height of trees and concentration of 1985, Wyant et al. 1986). The results of the study
foliage in the upper 50% of the bole. That reduces by Ryan et al. (1988) also confirmed observations
the probability of crown fires because sufficient by Bevins (1980) and Wyant et al. (1986) that tree
heat cannot easily reach the canopy to ig- nite the size, expressed by dbh, is inversely related to
crowns. Thick bark and extensive root mass in mortality.
mineral soil are other characteristics that make Peterson and Arbaugh (1989) evaluated factors
old-growth Douglas-fir fire-resistant. Saplings are related to the survival of coastal Douglas-fir 2
vulnerable to damage by surface fires because of years after wildfires in the spring of 1982 at four
their thin bark, resin blisters, closely spaced flam- sites in the western Cascades of Washington and
mable needles, thin twigs, and bud scales (Bradley Oregon. They used data collected on 294 trees
et al. 1992). with dbh >13 cm to develop a model for
Ryan et al. (1988) modeled long-term mortality
estimating post-fire survival. Their model showed
of Douglas-fir associated with damage to
that including both crown and bole damage
cambium and crown. The authors used data from
greatly improved estimates of post-fire survival.
the study
The authors pointed out that “rap- idly spreading
90
surface fires cause relatively greater amounts of
crown damage while ground fires with a long
80 Average diameter (cm)
duration of burning have relatively greater
Top logs
Mid-bole logs Butt logs potential for bole damage.” Therefore, including
70
30.48 variables that measure both crown and bole dam-
60 age provides greater latitude in estimating damage
for different types of fire. The authors emphasized
Log volume (%)

50
43.18 that their study identified factors related to post-
fire survival of coastal Douglas-fir, but did not
40 address the physiological effects of fire-caused
63.50
injury; as they stated, “carbohydrate production
30 and allocation are clearly important to the survival
of damaged trees but the relative impact of crown
20 and bole injury on these processes is unknown.”

10
Deterioration of fire-killed Douglas-fir
Wallis et al. (1974) determined the percentage of
2 3 4 5 total log volume of mature Douglas-fir decayed
Years after fire 2.5 to 5 years after fire-kill (Figure 12.3).
Figure 12.3 Percentage of total log volume of mature Douglas- Lowell et al. (1992) reviewed the literature on
fir decayed 2.5 to 5 years after fire-kill (from Wallis et al. 1974). the rate of deterioration of fire-killed and fire-
damaged
Chapter 12. Adverse Abiotic Factors 271

Douglas-fir. According to their review, factors Carlson (1980) documented damage to more
that influence the rate of decay include local site than 2,000 ha of interior Douglas-fir by a sulfate
con- ditions, aspect, elevation, slope, and soil, pulp and paper mill in western Montana. Damage
which influence deterioration by affecting the ranged from defoliation to various degrees of
temperature and moisture of a site. Additionally, discoloration and necrosis of foliage. Histological
these factors, coupled with precipitation, can lead study of needles showed that green-yellow color
to different rates of deterioration: indicated initial breakdown of chloroplasts and
• Diameter and age of tree. A large-diameter tree plasmolysis of me- sophyll cells. Yellow color
generally will deteriorate more slowly than reflected some collapse of mesophyll cells,
a small-diameter tree, and a young tree hypertrophy of phloem elements and parenchyma,
will deteriorate more rapidly than an and partial collapse of albuminous cells and
older tree. endodermis. Necrosis indicated collapse of
• Severity of burn. Less severely burned trees mesophyll, endodermis, and albuminous cells.
on a moist site tend to deteriorate more slowly Carlson and Gilligan (1983) showed in laboratory
than those severely burned. The opposite is and field studies that phytotoxic gases caused
true on dry sites where severely burned trees histological symptoms in Douglas-fir needles
take longer to deteriorate because of lack of
distinct from those induced by winter drying,
moisture.
normal drought, or salt. They emphasized that the
The thin sapwood of Douglas-fir has mostly dete- “identification of injury and related forest
riorated by the end of the third year post-fire; in damage near sources of phyto- toxic air
the fourth year, the heartwood will begin to emissions may be confounded by insects,
deteriorate. The heartwood of Douglas-fir is diseases, weather or other abiotic factors.” Their
moderately durable: coastal Douglas-fir of 60 to study demonstrated that histological procedures
250 years old (diameter range: 53–76 cm) takes 3 can be very helpful for the correct identification
to 4 years to reach 50% deterioration; trees 200 to of foliar chlorosis and necrosis. Leininger et al.
400 years old (diameter range: 51–60 cm) take 10 (1991) ranked seedlings of Douglas-fir as the
to 15 years to reach 50% deterioration; trees 400 most sensitive of five coniferous species to
years and older (diameter range: 130–152 cm) simulated ambient SO2 exposure. Their ranking
require 20 years to deteriorate by 50%. was essentially the same as that by Scheffer and
Hedgecock (1955) and by Carlson (1980) in field
Air Pollutants studies of older Douglas-fir.
Sulfur dioxide Emissions of SO2 caused annual increment
losses of more than 50,000 m3 in the forests of
Scheffer and Hedgecock (1955) wrote that SO2
eastern Germany since the 1960s. To explore the
dam- age to Douglas-fir had been reported as
possibility of breeding for SO2 resistance, a
early as 1912. The SO 2 emissions came from the
program was begun in the former German
Washoe smelter close to Anaconda, Montana.
Democratic Republic to in- vestigate the
Douglas-firs were dying as far as 8–13 km in all
heritability of SO2 resistance in several conifers.
directions from the smelter. Year-ring analyses
Heritability was shown to be relatively high (h 2 =
showed a distinct decrease of radial growth from
0.6) in Douglas-fir (Tzschacksch 1981).
1892 through 1910. Scheffer and Hedgecock
Surprisingly, coastal Douglas-fir showed greater
(1955) investigated SO2 inju- ry to coniferous
SO2 resistance than did the more frost-hardy
forests in the upper Columbia River Valley from
interior Douglas-fir (Tzschacksch 1982).
smelters near Kettle Falls, Washington, and Trail,
British Columbia. Douglas-fir was among the Fluorides
species damaged or killed. They stated that in- Fluorides are among the air pollutants known to
juries caused by frost and drought resembled SO 2 damage conifers, including Douglas-fir. Estimates
injury in some respects, but were not responsible are that about half of fluoride emissions from in-
for the damage in the upper Columbia River dustrial processes (Semrau 1957) are gaseous and
valley.
half are particulate. Fluorides enter needles
mainly through stomata. Once in the foliar
tissue, they are in
272 Douglas-fir: The Genus Pseudotsuga
a soluble state and tend to accumulate at needle m3, approximately equal to the annual cut
tips, causing necrosis. Treshow et al. (1967)
documented growth decline and mortality of
Douglas-fir near a phosphate reduction plant in
Idaho. They found up to 100% reduced
diameter growth when the foliar fluoride
concentrations exceeded 50 ppm. Foliar
fluoride levels in excess of 100–200 ppm
caused mortality, but precise threshold levels
could not be established. A study initiated by
the USDA Forest Service in 1969 (Carlson and
Dewey 1971) showed that fluoride emissions
from an aluminum reduc- tion plant in
northwestern Montana caused varying degrees
of visible fluoride injury to vegetation on
28,000 ha in parts of the Flathead National
Forest and the southwestern portion of Glacier
National Park. Elevated fluoride levels were
found in vegeta- tion on nearly 86,700 ha of
forested lands of mixed ownerships.
Conifers that showed tissue necrosis and
elevated fluoride levels were ponderosa pine,
lodgepole pine, western white pine, and Douglas-
fir. Trees differed in susceptibility to fluoride
injury shown by visual burn symptoms. White
pine was most susceptible, followed by ponderosa
pine, lodgepole pine, and Douglas-fir. Carlson et
al. (1979) conducted an in- tensive field study in
an attempt to relate visible foliar degradation to
foliar fluoride accumulation. They sampled
110,000 needles of Douglas-fir, west- ern white
pine, and lodgepole pine. Each sample consisted
of needles that had formed in 1975, 1976, and
1977. Foliar injury generally appeared at fo- liar
fluoride concentrations of less than 10 ppm.
Mottling or chlorosis of foliage was evident at less
than 6–8 ppm in Douglas-fir. Apparently, a
threshold effect is absent; adverse effects began at
slightly over baseline concentrations.

Wind
Wind can do serious damage to Douglas-fir by
break- age of trunks and blowdown. That was
spectacularly demonstrated by the Columbus Day
windstorm on October 12, 1962, which caused
more damage to the forests of the Pacific
Northwest than any other windstorm in recorded
history. The blowdown of timber, most of it
Douglas-fir in western Oregon and western
Washington, amounted to more than 26 million
in the two states at that time (Lynott and “breeding would probably be complex, expen-
Cramer 1966). But less violent windstorms sive, and slow. In contrast to breeding,
have also caused considerable damage to silvicultural techniques to minimize windthrow
Douglas-fir stands within and outside its may be simpler and less expensive.”
natural range. Among factors con- tributing to
damage may have been cutting practices
(Ruth and Yoder 1953, Munger 1954,
Gratkowski 1956), topography (Steinbrenner
and Gessel 1956), waterlogged soils (Prior
1959), root structure and penetration (Groth
1927, 1928; Soest 1954), crown size and
shape (Brown and Jones 1989), stem and
wood characteristics (Studholme 1995).
Douglas-fir ranks higher as a windfirm
species than most commercially important
conifers (Henkel 1960, Brünig 1974,
DeChamps et al. 1982). This rank- ing is based
mainly on observations made after storm
events. An exception is a study by Moore and
Gardiner (2001), who investigated the effect
of silvicultural practices on the relative
stability of Douglas-fir and Pinus radiata by
calculating the critical wind speeds for
damage at yearly intervals over the length of
typical rotations of each species. They
concluded that “a Pinus radiata stand grown
on a 28-year rotation was three times more
likely to suffer catastrophic wind damage than
a Douglas-fir stand grown on a 45-year
rotation. The most critical factor behind these
differences was the lower drag coefficient of
Douglas-fir foliage.”
The genetic test site of the Pacific Forest Research
Station of the USDA Forest Service in the
Willamette Valley was in the main path of the
wind storm of January 7, 1990, a storm of an
intensity expected once in 20-25 years (Silen
et al. 1993). The damage caused by that storm
provided a rare opportunity for an
investigation to estimate the genetic
component in susceptibility to blow-down
among F2 full-sib families of 6- and 7-year-
old coastal Douglas-fir. The results indicated
that susceptibility to windthrow differed by
family and was significantly related to their
height, but that height accounted for less than
a third of the genetic component of variation.
The authors concluded that their findings
“suggested the possibility of successful
breeding for resistance to windthrow in
Douglas-fir,” but they also cautioned that
13. Ontogeny
Denis P. Lavender

A
ccording to Bond (2000), “woody perennials 10-,
do not appear to go through a defined senes- 20-,
cence phase but do have predictable devel- 40-
opmental stages,” i.e., juvenility, maturity, and old and
age. We agree and in this section, we will review
the morphological, anatomical, and physiological
changes that occur as Douglas-fir develops from a
seedling to the tree we call “old growth.” We will
also discuss environmental and other factors such
as insects and diseases that can affect Douglas-fir
at
different life stages.

Ontogeny – Growth
Foliage, anatomy, quantity, distribution
Perhaps the earliest report of leaf maturity chang-
ing with age is that of Goldfarb et al. (1991), who
reported that the development of buds on
cotyledons in response to applications of
cytokinins decreased with cotyledon age. Ritchie
and Keely (1984) found that needle weight
declined with seedling age (from 1 to 9 years).
Working at the Wind River Canopy Crane site
in Washington and in the Cascade Mountains of
Oregon, Apple et al. (2002) found that the
anatomy of Douglas-fir needles “differed
significantly be- tween old-growth trees and
saplings at all sites, suggesting a developmental
change in needle anat- omy with increasing tree
age,” noting that “needles of saplings were longer
and had proportionately smaller vascular
cylinders, larger resin canals and few hypodermal
cells“ (p. 129). They also found that “needles of
old-growth trees had an average of 11% less
photosynthetic mesophyll area than needles of
saplings. The percentage of non-photosynthetic
area in needles increased significantly with
increasing tree age from the chronosequence of 273
450-year-old trees at the Wind River site” (p.
129). Apple et al. (2002) speculated that the
reduction in photosynthetic area in older trees
may contribute to their decreased growth rates.
Meinzer et al. (2008) suggested that tension af-
fected needle growth of tall trees. (Overton et al.
1973) found that the N content of old-growth
foliage (0.96%), but not that of other elements,
was lower than that commonly reported for young
growth (1.42%).
Seedlings
Ritchie and Keeley (1994) studied Douglas-fir at
ages 1 to 9 years and noted that needle weight
appeared to decrease with age as did
“Chlorophyll, chloro- phylls and total chlorophyte
concentrations which declined between 5 and
10% with aging between 1
and 9 years.”
Mature trees
Meinzer et al. (2008) noted that foliage growth
de- creases with height. Silver (1962) found that
28% of the foliage of a 50-year-old tree was
current and that 60% of the foliage in the upper
third of the crown was current, as opposed to 50%
of the lower third. Foliage of mature Douglas-fir
trees is a transi- tion type between juvenile and
old growth foliage. Maguire and Bennett (1986)
and Maguire and Batista (1996) reported that tree
dimensions are a good estimate of foliage
quantity.
Branches
Saplings
Ritchie and Keely (1994) found that as young
Douglas-fir plants matured, the most consistent
change was that nodal branches were shorter.
Frothingham (1909) observed that “sapling
Douglas-
274 Douglas-fir: The Genus Pseudotsuga

fir had long slender branches at relatively wide Saplings


intervals on the trunk” (p. 9).
The largest height increment occurs between 20
Old growth
and 30 years, and the ability to maintain a fairly
Bond (2000, p. 349) noted that “the production rapid height growth is maintained over a long
of new primary branches generally stops when period.
maxi- mum height is achieved and branch
extension also slows down. This is when the Mature trees
characteristics of old growth emerge. Leaf Hermann and Lavender (1990) noted that
bearing stems tend to be thicker and the leaves Douglas-fir in high elevation forests of the
themselves are often thicker and smaller on old Oregon-Washington Cascade range can continue
growth compared with leaves on young mature height growth at a sub- stantial rate for more than
trees.” 200 years. Frothingham (1909) recorded similar
Hummel (2009) reviewed several papers and data for trees in the interior. Hermann and
cited previously unpublished material finding that Lavender (1990) summarized height growth
branch size of trees in Douglas-fir forests of patterns for older Douglas-fir as follows:
Oregon, Washington, and California varied with
Height growth of Douglas-fir on dry sites at mid-side
crown and tree parameters; for physically indices in the Cascade Range of western Oregon is
comparable trees, branches on old-growth had similar to that of upper-slope Douglas-fir in the
greater diameters, prob- ably because lack of Washington and Oregon Cascade Range. At higher
site indices, however, height growth on dry sites is
elongation focuses growth on diameter.
initially faster but slower later in life; at lower site
indices, it is initially slower but faster later in life.
Ontogeny – Height On a medium site (III) at low elevations, height
growth, which averages 61 cm (24 in) annually at age
Seedlings
30, continues at a rate of 15 cm (6 in) per year at age
Douglas-fir in western Oregon, Washington, and 100, and 9 cm (3.6 in) at age 120 (18,39). Trees 150
British Columbia may be among the tallest trees to 180 cm (60 to 72 in) in diameter and 76 m (250 ft)
in height are com- mon in old-growth forests (22).
in the world (Hermann and Lavender 1990), but The tallest tree on record, found near Little Rock,
early seedling height growth is slow for the first 5 WA, was 100.5 m (330 ft) tall and had a diameter of
years before it then begins to accelerate. Height 182 cm (71.6 in). Coastal Douglas-fir is very long
lived; ages in excess of 500 years are not uncommon
growth of seedlings is very much a function of
and some have exceeded 1,000 years. The oldest
seedling age. Early height growth of seedlings is Douglas-fir of which there is an authentic record
relatively slow (Williamson and Twobley 1983). stood about 48 km (30 mi) east of Mount Vernon,
Measurements in the nursery suggest that second- WA. It was slightly more than 1,400 years old when
cut (39).
year seedlings may grow 270 cm (Krueger and
The interior variety of Douglas-fir does not attain
Trappe 1967), whereas young trees on plantations the growth rates, dimensions, or age of the coastal
commonly reach a mean of 135 cm in 5 years. variety. Site class for Rocky Mountain Douglas-fir is
After 5 years, the growth is much faster. Ten-year- usually IV or V (Site index 24 to 37 m or 80 to 120 ft
at age 100) when compared with the growth of this
old seedlings are commonly at least twice the species in the Pacific Northwest. On low sites, growth
height of 5 year-olds. And saplings grown in is sometimes so slow that trees do not reach saw-log
favorable environments have produced leaders size before old age and decadence overtake them.
165 cm in length (Newton, personal Interior Douglas-fir reaches an average height of 30 to
37 m (100 to 120 ft) with a
communication). But perhaps the best measure of d.b.h. between 38 and 102 cm (15 and 40 in) in 200 or
the effect of the environment on seedling growth 300 years. On the best sites, dominant trees may attain
is the 2-year-old seedling in a greenhouse under a height of 49 m (160 ft) and a d.b.h. of 152 cm (60
in). Diameter growth becomes extremely slow and
continuous long photoperiods, which measured
height growth practically ceases after age 200. Interior
300 cm. Campbell (1972) and Overton and Ching Douglas- fir, however, appears capable of response to
(1948) both reported that the environment has a release by accelerated diameter growth at any size or
greater effect on seedling height growth than does age. The interior variety is not as long lived as the
coastal variety and rarely lives more than 400 years,
genetics or age. although more than 700 annual rings have been
counted on stumps. (Hermann and Lavender 1990, p. 534)
Chapter 13. Ontogeny 275

Old growth vations of Hermann and Lavender: “Biomass accu-


Bond et al. (2007, p. 441) presented a detailed in general for height growth that agreed with obser-
study designed to separate size and age effects
on tree height and growth. They found that,
On high quality sites, maximum height growth of
Douglas-fir can exceed 1.5 m year –1, and trees may
achieve heights greater than 75 m, whereas maximum
height growth and total maximum height of trees with
similar genetic potential on poor sites can be a small
frac- tion of these values Clearly, height growth is
strongly
influenced by environmental conditions. However,
new insights emerge when height growth is viewed as
a function of height . . . .
Under all site conditions, the maximum rate of
height growth of trees occurs while they are relatively
small; subsequently, growth declines as a linear
function of height for more than a century. [Douglas-
fir] trees
lose, on average, about 2 cm year–1 in height growth
for each new meter of growth irrespective of site
conditions after they reach their growth maximum.
(Bond et al. 2007, pp. 441–442)

They examined the evidence for factors causing


the above and concluded that “size, not age,
drives developmental changes in height growth in
Douglas- fir. Reduced carbon assimilation does
not play an important role in height growth
decline” (p. 441). They also noted: “We found
that neither intrinsic aging nor photosynthetic
reduction due to hydrau- lic constraints or other
factors is a likely cause of DDHG” (p. 451) or
“developmental decline in height growth.”
Ryan and Yoder (1997) examined several hy-
potheses that attempt to explain why trees decline
in height growth with age, after a maximum when
relatively young: i.e., respiration, nutrient
limitation, genetic changes in meristem tissue, and
hydraulic limitation, concluding that the last is
most likely. They supported this conclusion by
noting factors that reduce the effects of each of
the first three hy- potheses, observing that
“stomata and consequently, transpiration and
photosynthesis is most respond- ing to changes in
hydraulic resistance. Hydraulic resistance must
increase with tree height or tree age.
Photosynthesis must be lower in the foliage of
older trees” (p. 239). They present evidence
supporting each of the above. Domec et al. (2006)
concurred that hydraulic resistance increases with
height.
Ryan et al. (1997) presented data for forest trees
mulation and growth of even-aged forests A number of changes occur in seedlings between
follows a universal pattern as the trees the germinant stage and 15 years. One major
increase in size, growth is slow initially, change is the development of reproductive
increases as leaf area develops, peaks as leaf capacity. Douglas- fir commonly initiates
area reaches its maximum, and then declines production of reproduc-
for the majority of the stands’ lifespan” (p.
215). They discussed in detail the following
possible reasons for growth decline: “(1)
changes in photosynthesis, (2) change in
nutrient supply, (3) change in respiration,
(4) change in fine root production and
longevity, (5) allocation to symbionts, (6)
allocation to foliage and branches, (7) change
in maturation” (p. 213). They concluded that
only the first reason (changes in
photosynthesis), as a result of reduced leaf
area or photosynthetic capacity, is likely (p.
251). Ryan and Yoder (1997, p. 241) asked
the same question with regard to maximum
tree height and concluded that hydraulic
limitation and not respiration, nutrient
limitation, or genetic change is most likely (p.
241)
A maximum height that varies with resource
availability and slower height growth in older
individuals appear to be universal for trees, old
trees are different both physiologically and
morphologically from younger trees. They have
lower rates of photosynthesis, reduced height
and diameter growth rates, and a distinctive
architecture. Nutrition, Carbon allocation
including respiration, meristematic activity,
and trees’ hydraulic architecture can all
potentially change with tree growth and
promote slower growth in older trees, in fact,
these processes may interact. (Ryan and Yoder
1997, p. 244)

Bond (2000) presented a summary of the


changes that occur with age in old woody
plants, with par- ticular emphasis on
photosynthesis. She noted:
Published studies from a variety of experimental
situ- ations generally indicate that both
photosynthesis and stomatal conductance are
reduced with the age of shrubs and trees. These
degrees have been reported at all phases of
development: seedlings versus older plants,
seedlings versus juveniles versus mature plants,
mature versus old growth. Might immobilize
Nitrogen in some
aging forests, making it less available for new
growth. Reduced photosynthesis is a likely
consequence because Nitrogen content of leaves
is closely correlated with photosynthetic
capacity. (Bond 2000, p. 350)

Ontogeny – Phenology
276 Douglas-fir: The Genus Pseudotsuga

tive buds at approximately 10–15 years (Isaac and • Bud set is strongly inherited.
Dimock 1960) and the production of cones
• Saplings are determinate.
increases until the tree is 275 years old or older.
• Bud set is in mid-June.
Other indica- tors of changes in juvenility are
“capacity of cuttings to root, seedling mainstem Emmingham (1997) noted that bud burst ap-
diameter, nodal branch length, diameter, all peared to be triggered after soil temperatures
increased with increasing age” (Ritchie and Keely reached 5ºC. Late wood formation generally
1994). Robinson and Wareing (1969) concluded began after 90% completion of leader growth.
that phase change occurs after the meristems have Under fa- vorable conditions growth cessation
undergone a number of divisions so that phase was long; the cambial growth continued to late
change is correlated, but not determined by October. Walters and Soos (1963, p. 83) noted
attainment of a certain size. Interestingly, phase that lower branches had a shorter elongation
change in old growth is a function of size, not age period than the leader but that laterals had a
per se. greater growth rate than the leader. Annual
growth is not correlated with bud burst date.
Seedlings
Saplings
Li and Adams (1993) found the following:
Emmingham (1977) discussed in detail the
• Seedlings are indeterminate; bud set is in early phenol- ogy of several Douglas-firs seed sources
fall. shown at several different locations, and with all
• Late buds are susceptible to summer drought, sources and locations from Oregon, generally
positively correlated with height. midway between the Columbia River and
• Early bud break is weakly correlated with California. His data showed that all seed sources
height. initiated both root and height growth in April in
• Bud burst phenology is under strong genetic the low elevation areas, but not until June at the
control. plantation at 1,050 m in elevation. Cambial
• Bud set is weakly inherited. growth started a few days after bud swell for most
White et al. (1979) noted that seedlings from seed sources. Growth appeared to be trig- gered
southerly or low-rain areas broke buds early. In after soil temperatures reached 5°C. He found that
France, Michaud and Najar (1980) found the “leader growth was nearly completed by the end
follow- ing for populations of seedlings of August at low land sites and by mid-August at
representing almost the entire natural range of the Cascade Mountain plantation (1050 m).
Douglas-fir: Drought was most severe in the Coast Range and
Corvallis planta- tions, where shoot growth
• Seedlings from east of the Cascades broke
buds stopped first. Latewood formation generally
early. began between 90% completion of leader growth
• “Latitude has an important influence on and 90% completion of cambial growth (p. 154).
flushing – late provenances are situated in a Emmingham (1977) suggested that the “cessation
zone bounded by latitude 44° to the south and of shoot growth while temperatures and moisture
latitude 49° to the north.......Within this zone conditions were favorable, was keyed to shorter
altitude has little influence on flushing” (p. day length” (p. 161). Cambial growth continued to
192). late October for all seed sources in all areas.
• An important relationship between vigour and Farther north, Walters and Soos (1963) examined
flushing date was observed, the most vigorous the phenology of Douglas-fir saplings on two
provenances are the least susceptible to late elevations, 100 m and 500 m, in southern British
frosts. Columbia.
• Sapling height, dbh, and bole volume are
correlated with budburst. Ontogeny – Photosynthesis
• Bud burst phenology is under moderate to It is very difficult to relate photosynthetic rates to
strong genetic control and is highly stable.
the age of the tree because these rates are a
function of
Chapter 13. Ontogeny 277

both the needles and the environment. Bond This is in strong contrast to the data reported by
(2000) noted in a detailed review “that published Silver (1962), Dice (1970), and M. Tohell
studies from a variety of experimental situations (personal communication), all of whom noted that
generally indicate that both the photosynthesis the major- ity of the foliage of young Douglas-fir
and stomal conductance are reduced with the age is less than 2 years old. Woodman (1971) noted
of trees and shrubs,” and that “these decreases that the photosyn- thesis rates in foliage of young
have been re- ported at all phases of development: trees was maximal for current needles; while 1-
seedlings vs. older plants; seedlings vs. juvenile year-old needles had photosynthetic rates 72% of
versus juveniles vs mature plants; juveniles vs maximum; 3-year-old needles, 50%. These
mature and or old growth” (p. 350). She noted differences complicate estimates of tree age effects.
reduced photosynthesis with reduced N, which is Apple et al. (2002) noted that old growth
closely correlated with photosynthetic capacity. needles are less efficient photosynthetically than
those of young growth. Thomas and Winner
Young trees
(2007) con- clude that “in general measure LMA
According to McArdle and Meyer (1930),
(leaf mass/ leaf area) ratio in old trees leads to a
The inability of Douglas-fir to live in its own dense decreased photosynthetic capacity. A number of
shade insures, in well-stocked stands, the early and
gradual shedding of the lower branches and the
studies have discussed changing photosynthesis
production of clean lumber thereafter. Sensitivity to with tree age (Ryan et al. 1997, Ryan and Yoder
shade varies 1997, Bond 2000, Thomas and Winner 2002,
with age. Young trees being more shade-tolerant than
Ryan et al. 2006, Bond et al. 2007). Parker (1997)
old trees. Likewise, trees having favorable growth
con- ditions are more tolerant of shade than those on noted the difficulty of measuring light intensity in
poorer sites. Early in life, Douglas-fir is able to old growth stands and Thomas and Winner (2002)
withstand some side shading, but after the 25th year
the tree is unable to make satisfactory growth or live
noted that LMA in ma- ture trees leads to
in either side or overhead shade. (p. 4) decrease in photosynthesis. Ryan et al. (1997)
Chen and Klinka (1997) found that shade grown reported lower photosynthesis in older trees,
Douglas-fir foliage had higher photosynthetic suggesting the following reasons:
rates based on dry weight at all light intensities • increased hydraulic resistance
than did open grown foliage for Pseudotsuga • reduced leaf area caused by crown abrasion
menziesii var. glauca seedlings. Lewis et al. (2000, • reduced nutrient supply
p. 454) suggested that “the high photosynthetic
• reproductive effort
capacity of Douglas-fir is consistent with its
• increased mortality of older trees
dominance of early successional environments.”
As Bond (2000) noted, however, “shaded plants • genetic changes with meristem age
in the forest understory (presumably older than Several studies noted that photosynthesis may
seedlings) generally have low photosyn- thetic be limited by various factors of the environment
capacity compared with sun-adapted plants.” This (Helms 1964, 1965; Woodman 1971; Parker
probably explains changing tolerance to shade in 1994). Helms (1965) noted that Douglas-fir may
Douglas-fir. A number of reports (Hodges 1967, photosyn- thesize at low light intensities. None of
Brix 1970, Del Rio and Berg 1979, Drew and these reports indicated differences with age.
Ferrell 1987, Chen et al. 1996, Emmingham 1997,
Competition
Khan et al. 2000) have suggested that Douglas-fir
seedlings tolerate some degree of shade, although Dawkins (2009) observes that competition results
Reed et al. (1983) reported that responses of tree in trees 30 m tall instead of 3 m with no
species to varying light availability differed with appreciable gain on the individual tree than
availability of nutrients and water. possible increased light absorption, and Darwin
(1860) noted that the most intense competition
Mature trees occurs between individu- als of the same species.
Overton et al. (1973) reported that 61% of the A number of reports (De Champs 1997; Timmis
foliage of old growth trees was more than two and Tanaka 1976; Van den
years old.
278 Douglas-fir: The Genus Pseudotsuga
Driessche 1984,a-d; Smith and Reukema 1986; increase damage on Douglas-fir:
Curtis and Reukema 1970), and unpublished
reports from Wind River, for ages from seedlings
through young second growth all noted that tree
diameter and height growth all increase with
lesser competition. Perhaps, the most striking
paper is that of Tappeiner et al. (1997), who found
that the large trees’ charac- teristics of old growth
forests developed as seedlings and young mature
stands of 100 trees per hectare, as opposed to
young stands today of 600 trees per hect- are and
that growth of the former was significantly greater
than that of current stands. Accordingly, large
trees in the future will develop only if current
stands are dramatically thinned. In addition,
Latham and Tappeiner (2002), who reviewed a
number of papers indicating that thinning
Douglas-fir stands increases tree height and
diameter, noted that old growth trees increase
growth after reduction of competition.
In contrast to the above, Scott et al. (1998)
found
that growth of 7- to 9-year-old Douglas-fir
seedlings increased growth with increased
stocking. This may be a function of light reflected
discussed earlier by Ritchie. A number of
references (Chappell et al. 1992, Gessel et al.
1979) and the numerous references cited therein,
noted that Douglas-fir is sensitive to nitrogen
fertilization throughout its lifespan, but noted
differences with age.

Ontogeny – Insects
Seedlings
Although are a number of insects that feed on
Douglas-fir seedlings, we will discuss only those
considered to have a major impact.
Conifer seedling weevil
The conifer seedling weevil (Steremnius carina-
tus) commonly feeds on bark near the groundline
(Condrashoff 1968). It is particularly damaging to
container seedlings, which typically have thin
bark. Adult weevils overwinter in the soil and
emerge in the spring; they are favored by warm,
moist periods. Weevils may live 3 years and are
favored by clearcut- ting, which produces many
favorable breeding sites, and by logging slash.
Burning, which destroys com- peting plants, may
1-year-old seedlings were more damaged than discussion of four insects that defoliate Douglas-fir:
2-0 seedlings. Condrashoff (1969, p. 2) found the western spruce budworm (Tortricidae:
as many as 17,500 weevils per hectare in Choristoneura fumifera-
cutover areas of British Columbia. Up to 10%
of seedlings may be killed by this weevil and
an equal number injured in the Pacific
Northwest.
Black army cutworm
In their detailed report, Ross and Ilnytzky
(1977) discussed the significant damage to
agricultural crops and to newly planted
conifers seedlings from the Black army
cutworm (Actebia fennica). The cut- worm
can cause 40% to 80% mortality on replanted
forest plantations that had previously been
burned. It overwinters as a first or second
instar or, possibly as eggs:
Moths fly late in summer and oviposit in the
soil, fre- quently on burned areas. Eggs hatch late
in the fall and the young larvae overwinter in the
soil. Feeding on sprouting vegetation begins
shortly after the snows recede in spring. Most
spring feeding in central B.C. in 1973-74-75
began after sunrise: night hours were frequently
too cool for larval activity. Feeding was com-
pleted in the valley bottoms about the second
week in June; it continued for a week or two
longer at higher altitudes.
There is some suggestion that a series of
warm, dry years precedes a black army cutworm
epidemic and that warm dry conditions during
egg laying and hatching are necessary for
development of infestations. Heaviest
concentrations of cutworm larvae have appeared
on clearings burned over the previous year or
two. . . .
Although few in number, the most severe or
exten- sive epidemics in North America have,
within a given region, occurred at about 20-year
intervals, lasting for 2, 3 and, in one instance, 4
years. The epidemic in 1973 covered more than
1,400 ha (3,400 acres), in 1974 about 2,800 ha
(7,000 acres) and in 1975 over 650 ha (1,600
acres). Logging and reforestation, including
prescribed burning, as practised in recent years,
may aggravate the duration and even the
frequency of epidemics. (Ross and Ilnytzky
1977, p. 4)
Mortality of seedlings was highest on dry,
burned sites (Shepherd et al. 1993).

Mature trees
Although defoliators are attack all ages of
Douglas- fir, they are perhaps most significant
in mature for- ests (primarily Pseudotsuga
menziesii var. glauca). The Defoliator
Management Guidebook (British Columbia
Ministry of Forests 1995b) gave a detailed
Chapter 13. Ontogeny 279
synchronous and dependent upon heat accumula-
na) and the Douglas-fir tussock moth
(Lymantriidae: Orgyia pseudotsugata), which are
major defoliators, and the western hemlock looper
(Geometridae: Lambdina fiscellaria lugubrosa) and
the western black- headed budworm (Tortricidae:
Acleris gloverana), which are lesser. The
guidebook is a source for much of the following.
Western spruce budworm
The western spruce budworm is discussed in
Chapter 9, so we do not cover it in detail here,
other than a brief comment on research
concerning potential factors involved in Douglas-
fir resistance. Clancy (1992 a,b) related
concentrations of mineral and organic compounds
to the incidence of spruce bud- worm and found
that susceptible trees had lower levels of sugars
than resistant trees. “No detectable difference in
foliar concentrations of terpenes, how- ever,
susceptible trees had a greater proportion of
monoterpenes, whereas resistant trees had greater
proportion of oxygenated monoterpenes. Resistant
trees also had delayed budbreak and growth com-
pared to susceptible trees” (see also Clancy et al.
2004, Brookes et al. 1985).
Douglas-fir tussock moth
“While there are thousands of insect species
native to British Columbia, very few of them
cause seri- ous concern to foresters. One of
these select few is a small caterpillar with a
characteristic coat of rust coloured tufts and
orange markings, known as the Douglas-fir
Tussock moth, Orgyia pseudotsugata”
(Anonymous 1990, p. 1). Beckwith (1978)
discussed the physiology and ecology of this
insect in detail, noting the following: “Number
of instars and body color vary depending on
factors such as genetic coding, sex, food quality
and quantity, temperature, and population
density. The females usually have one more
instar than males” (p. 27). There are five to
seven instars; “the final instar spins a grayish
brown spindle-shaped silken cocoon, which
incorporates some of the larval hairs, and is
spun on foliage, branches, and boles of host
trees” (p. 28). “If not disturbed, each female
lays all her eggs in a single mass,” which
generally contains 150 to 200 eggs (p. 30). Egg
hatch and tree bud break are generally
moths emerge in late summer, and about 2 weeks after
tion (C° > −14.7). The only acceptable food is pupating. The wingless adult females remain on the
cocoon, where they mate with
new foliage from buds. Larvae concentrate on
the winged male. The cycle is complete when the
the top of tree. Most dispersal is over short female deposits a new egg mass on or near her
distances. Larval development is related to cocoon. (Brookes et al. 1978, p. 191)
temperatures with alternat- ing day and night
lows. Mating and egg laying take place on the
cocoon, generally the day the female emerges.
The tussock moth may kill up to 40% of the
trees in an infected stand. Further, “Douglas-
fir trees that have been weakened by the
tussock moth defoliation may also be
susceptible to attack by other insect pests,
such as the Douglas-fir beetle” (British
Columbia Ministry of Forests, Lands, and
Natural Resource Operations 2014). Thus,
even light defolia- tion by the tussock moth
may be responsible for the tree’s untimely
death.
Brookes et al. (1978) discussed the life history of
the tussock moth as follows:
The Douglas-fir tussock moth is a univoltine
insect that overwinters in egg masses containing
150–250 eggs, usually on the underside of small
branches. Eggs
hatch in the late spring about the time that buds
break and new shoots begin to expand. These
events are closely synchronized, giving new
larvae the food they need for growth and
development. Shoot elongation
progresses rapidly in June and July, resulting
in a sup- ply of new foliage for the larvae to
feed upon. Neonate larvae normally leave their
egg masses at the time shoot elongation is
about 50% completed; by the third instar, shoot
growth has usually terminated. Small larvae
produce silk strands on which they can be dispersed by
wind throughout the stand.

According to Gast et al. (1991), larvae on a


host tree then “crawl to the top or to the ends
of brances and feed on the new foliage. The
first two instars of tussock moth feed
exclusively on the underside of the succulent,
new needles. This feeding damages the
needles, causing them to dry and turn red-
brown by midsummer. Faded foliage at the
tops of trees is often the first sign of tussock
moth infestation“ (pp. 19–20). Brookes et al.
(1978) detail the following:
After the new foliage has been destroyed, later
instars feed on old foliage; during an outbreak,
large portions or even the entire crown may be
defoliated. The five to six larval instars feed for
about 60 days and then spin
cocoons in foliated portions of the crown,
sometimes in crevices of the bark, or on dead
branches and twigs in the lower crown. The
280 Douglas-fir: The Genus Pseudotsuga
The tussock moth is found in the drier part of the may be instrumental in initiating decomposition and
Douglas-fir range, this it attacks only var. glauca. perhaps in determin-
Webb (1978) noted that defoliation of the up-
per crown by the tussock moth resulted in higher
light intensity received by remaining needles and
the photosynthetic efficiency of the tree is raised;
soon, increases may not occur until the needles
are accustomed to higher light levels. Hence, the
long- term loss of photosynthesis by defoliation is
not proportional to the quantity of foliage
removed. This is accompanied by a significant
drop in requirements by the remaining foliage.
The tussock moth preferen- tially feeds on new
foliage, which has, perhaps, the greatest
photosynthetic capacity. In contrast, when the
insect takes older foliage, that not only reduces
photosynthetic capacity, but food.
Western hemlock looper
The western hemlock looper (Lambdina
fiscellaria lugubrosa) “primarily attacks
hemlock,” but during outbreaks, “the looper
feeds on almost any foliage, including broad
leafed forest trees and shrubs” (British
Columbia Ministry of Forests 1995b). It is
“periodically destructive in coastal and interior
forests, reaching outbreak proportions every 11
and 20-plus years, respectively. Outbreaks of the
hemlock looper usually last about 3 years, after
which they are generally brought under control
by the action of parasites, predators, and
diseases. Heavy rains during the moth flight
period can reduce egg-laying and hasten the
decline of an outbreak.” From May through
early July, feeding by early instars is slight.
From mid-July until October, the larvae “feed
vora- ciously on both new and old foliage.”
Feeding starts in crown tops, later, more of the
crown is affected (British Columbia Ministry of
Forests 1995b, p. 58). (McCloskey et al. 2009)
noted that if global warming increases
temperature and reduces moisture, it has the
potential to increase hemlock moisture 100%.
Douglas-fir beetle
Zhong and Showalter (1989, p. 941) noted that
each beetle species or functional group could be
expected to inoculate and colonize boles with a
distinct micro- flora and that patterns of wood
utilization by various beetle functional groups
ing long-term rates of decomposition in boles spring through mid- August. It broods overwinter
of different tree species. within phloem in the adult, pupal, or late larval
The Douglas-fir beetle (Dendroctonus stages. Stark (1965) noted that fertilization
pseudotsugae) differs from the preceding reduces bark beetles (p. 800). Ross
insects in that it attacks the bole, not the
foliage. Ross et al. (2006) reported the
following:
The Douglas-fir beetle is normally present in
forests at low densities, breeding in Douglas-fir
trees that are injured or have recently died.
Tunneling by adults and larvae beneath the bark
produces a characteristic pattern distinguishing
the Douglas-fir beetle from other bark beetles
Beetle larvae need fresh, moist phloem (inner
bark) for food, so trees that have been dead for
more than a year are not suitable habitat. Injured
or recently killed trees have little or no defensive
capability, making them ideal sites for beetle
larvae to feed and develop. Periodically, natural
or human-caused disturbances such as
windstorms, fire, defoliator outbreaks, or
logging, create an abundance of suitable breeding
sites that al- low the beetle populations, under
favorable conditions, to rapidly increase to high
densities. At high densities, beetles are forced to
attack healthy, live trees because there are not
enough stressed and dead ones to support the
population By attacking a live tree in large
numbers, the beetles are able to overcome the
tree’s natural de- fenses and successfully
reproduce. (Ross et al. 2006, p. 5)

Global warming will create such conditions


and result in massive beetle kill of previously
healthy trees. Douglas-fir beetles
preferentially attack large, old trees in dense
stands with a high Douglas-fir component. As
Ross and Daterman (1997) noted, the
Douglas-fir beetle “usually breeds in portions
of tree boles that are greater than about 20 cm
in diameter. At low population densities, most
infestations occur in trees that have recently
died or live trees with lim- ited defenses
resulting from stress or injury. When
populations reach high densities, large
numbers of healthy trees may be successfully
attacked and killed” (Ross and Daterman
1997, p. 135). According to Ross and
Daterman (1995), the beetle “is found
throughout the range of Douglas-fir
(Pseudotsuga menziesii) in western North
America [and] when
suitable breeding material is abundant and
weather conditions are favorable, beetle
populations can reach high densities causing
considerable tree mor- tality” (p. 805).
Douglas-fir beetles have only one
generation per year. Adults fly from early
Chapter 13. Ontogeny 281

and Daterman (1997) suggested that thinning may many ecosystems. It does not destroy entire
reduce bark beetle infection and that sanitation stands over large areas. The fungus does not
and salvage logging may reduce breeding sites. spread from spores, but from root contact; this is
why eliminating infected stumps is a method of
Ontogeny – Disease control. Thies and Sturrock (1995) estimated that
A distinct group of fungal pathogens attacks P. weirii occurred on 84% of commercial
Douglas-fir at each stage, from seedlings through forestland in the northwestern United States and
old growth. We shall confine our observations to caused a loss in wood volume of 40% to 60% in
those diseases deemed most damaging. areas affected. Nelson and Sturrock (1993)
noted that grand fir and Douglas-fir are the
Seedlings most susceptible of Pacific Northwest conifers
We have discussed nursery diseases previously. to
Natural reproduction is remarkably free of P. weirii.
disease, primarily because forest soils Van der Kamp (1993b) found that “young trees
frequently contain organisms that inhibit those are killed quickly (1-3 years from first symptoms
causing disease, i.e., damping off, Fusarium. to death); older trees (40–60 years old when first
Peterson (2008), in a lengthy review of infected) may survive for decades, but such trees
Fusarium, noted, however, that mishan- dled produce little increment, instead the host spends
seedlings may develop Fusarium after planting much of the available energy producing new roots
(Linderman 2000). to replace those killed by the pathogen” (p. 5).
An exception is Phellinus, which is spread by Bloomberg and Reynolds (1982) reported that
root contact and is not soilborne. Phellinus weirii, transferred mycelium of Phellinus weirii was
laminated root rot (Thies and Sturrock 1995), is endo- trophic rather than ectotrophic. They found
one of several pathogens particularly destructive that root diameter, rather than root depth, was
to Douglas-fir. Others are Phaeocryptopus positively correlated with infection.
gaeumannii, Swiss needle cast (Hansen et al.
2000), and Armillaria solidipes, formerly Saplings, young growth
Armillaria ostoyae (Van der Kamp 1994, Filip and Armillaria root
Ganio 2004). Armillaria root disease is an important disease of
both conifers and hardwoods throughout the
Young trees
world (Mallett 1992). There are currently 36
Phellinus weirii described spe- cies in the world (includng 10 in
Wallis (1976) stated that “Phellinus weirii North America). Blenis (1995) observed that
(Murr.) Gylbertson is the most destructive “Douglas-fir is very susceptible to Armillaria
disease in young growth Douglas-fir in British ostoyae. This fungus causes foliar discoloration,
Columbia” (p. 3). Thies and Sturrock (1995) resinous and a light yellow stringy decay of the
summarized research concern- ing P. weirii, wood, but the most positive symptom is a
noting that it is widespread in western Canada mycelial fan.” Blenis (1995) also noted, however,
and the northwestern United States. There are that Armillaria rarely kills trees over the age of
two fairly distinct forms of the fungus. One is a 20; Van der Kamp (1994) agreed.
common cause of root rot in western redce- dar In a series of papers, Entry et al. (1990; 1991
in the northern Rockies; the other commonly a,b; 1992 a,b; 1993) related the density of the
kills Douglas-fir and several other conifer Armillaria attack to the level of carbohydrates and
species throughout northwestern North defensive phenols in seedling roots. Redfern
America. It forms red ring rot, and is the most (1978) reported that supressed trees were least
destructive disease in the area. Trees weakened resistant, but that Douglas-fir is relatively resistant
by this disease are often killed by bark beetles. in Britain. Filip and Ganio (2004) found that
The pathogen is believed to have evolved with Armillaria ostoyae is the most common root
its host and is a natural, perhaps necessary part of disease in Douglas-fir planta- tions in Oregon.
282 Douglas-fir: The Genus Pseudotsuga

Mature trees, old particular stand” (p. 16). The distribution of P.


growth weirii may be either diffuse or aggregated. Minor
infections of both above diseases have been
Root and butt rots
reported under
Polyporus schweinitzii (Schweinitzii root and butt of either the presence or intensity of laminated root rot
rot), in a
Fomes pini (conk rot), Fomitopsis officinalis
(brown trunk rot or quinine fungus), and Fomitopsis
cajanderi (yellow brown top rot, formerly called
Fomes sub- roseus) all attack old-growth
Douglas-fir. Nelson et al. (1981) noted the “role
of P. weirii as a perennial inhabitant of the site,
substantially reducing produc- tivity, surpassing
importance as a killer of individual trees” (p. 1).
Thomson et al. (1996) found that P. weirii
reduced levels of chlorophyll a and nitrogen in
foliage. Thies (1983) found that P. weirii reduced
growth of Douglas-fir as much as 32%. Growth
loss was not related to crown symptoms or
number and size of infected roots.

Young stands

Phellinus weirii
Sturrock and Garbutt (1994) discussed the spread
of Phellinus weirii and its characteristics in the
two passages below:
Infection by P. weirii starts when healthy roots of sus-
ceptible tree species contact infected roots of an
adjacent tree or infected stumps and roots (residual
inoculum sources) from the previous stand. Surface
(ectotrophic) mycelium spreads from infected roots
onto the surface of healthy roots. Ectotrophic
mycelium eventually pen- etrates to the interior of
host roots, likely gaining entry through both intact and
injured bark. Once inside the root (as endotrophic
mycelium), the fungus progres- sively destroys root
tissue, depriving the tree of water and nutrients and
weakening its structural support. (Sturrock and
Garbutt 1994, p. 2)
Phellinus weirii can be positively identified in living
sus- pect trees by examining the root collar and lateral
roots for grey-white to tawny to mauve colored
ectotrophic mycelium. Brown crustlike mycelial mats
commonly form over surface mycelium below the
duff layer, par- ticularly in the crotches of roots.
Reddish brown hair- like structures called setal
hyphae may be seen with the aid of a hand lens,
scattered in surface mycelium or in pieces of wood
with advanced decay. Setal hyphae, in conjunction
with these other signs, are diagnostic of P. weirii.
(Sturrock and Garbutt 1994, p. 5)

According to Thies and Sturrock (1995), “as yet


no strong evidence exists that any individual site
factor, or group of factors, is a reliable predictor
laboratory conditions. Tricoderma viride was
shown to be inhibitory to P. weirii, but we
have found no reports of how significant such
effects are under field conditions (Goldfarb et
al. 1989 a,b).

Swiss needle cast


In a survey of the effects of Swiss needle cast
(Phaeocryptopus gaeumannii), in western
Oregon. Weiskittel et al. (2006) noted the
following:
The condition of young conifer plantations in
Europe is often rated by foliage retention,
alternatively defined as needle longevity or the
average number of needle cohorts held by the
trees (Innes 1993). Foliage retention is currently
the primary index of Swiss needle cast (SNC)
severity in Douglas-fir (Pseudotsuga menziesii
(Mirb.) Franco; Hansen et al. 2000). Other
foliage attributes such as crown color and crown
density have been explored as indices of SNC
severity, but foliage retention is less subjective
and performs better than or as well as other al-
ternatives for predicting tree and stand growth
(Maguire et al. 2002). SNC is caused by an
endemic fungal patho- gen, Phaeocryptopus
gaeumannii, whose hyphae grow into needles
through stomates and interrupt gas exchange by
occluding stomates with fruiting bodies, or
pseudothecia (Hansen et al. 2000). The fungus
causes premature loss of older foliage, reducing
mean foliage retention to as little as 1 year and
volume growth by as much as 50% (Maguire et
al. 2002). The disease eventually changes several
crown structural and morphological attributes
such as live crown length, branch size, and
specific leaf area (Weiskittel 2003). Currently,
over 72 000 ha in the Oregon Coast Range are
showing symptoms detectable by aerial survey,
reflecting the dramatic increase in this disease
since 1990 (Kanaskie et al. 2004). (Weiskittel et
al. 2006, p. 1498)

They also noted that “with increasing SNC


severity, the modes of the youngest three age-
classes shifted upwards, while the modes for
4- and ≥5-year-old foliage were located lower
in the crown relative to that of healthy trees”
(p. 1506).
Swiss needle cast differs from most
diseases of older Douglas-fir in that it is a
foliage disease, not a rot of bole or roots, and
that significant damage in western Oregon has
been reported only during the last two
decades. Losses as high as 25% in top height
growth, 49% in volume, and 48% in basal
area have been reported (Maguire et al. 2002).
Major losses to Swiss needle cast extend to a
band within about 30 miles of the coast (the fog
Chapter 13. Ontogeny 283
living trees” (p. 5). The populations of old
zone). Measurements of weather data and disease
virulence show a strong correlation between
maxi- mum temperatures in the period from
November to February (Manter et al. 2005) and
incidence of disease severity. This suggests the
possibility that the disease is favored by a lack of
chilling of saplings and represents an early effect
of global warming.
Weiskittel and Maguire (2007) noted:
“Defoliation from the endemic SNC pathogen
can drastically reduce LAI (leaf area index) and
change both total and seasonal foliage litter fall
patterns” (p. 121). Hansen et al. (2000) noted
that many trees had high defoliation in the
upper crown. They also reported that “fungicide
applications reduced pseudothe- cial density
and increased needle retention. Most infection
occurs in newly emerged current season
needles” Hansen et al. (2000, p. 775).

Mature stands, old growth


Swiss needle cast
Little research has concerned Swiss needle cast
in stands of older trees, as Shaw et al. (2011) also
noted. It has been found on old trees, however,
and the symptoms follow the same pattern as
found in young stands. Black et al. (2010),
working in mature stands in the Oregon Coast
Range, reported that Swiss needle cast is an
increasing threat to old stands, and also noted
effects from warmer temperatures: They found
that “even mature forests of natural origin are
susceptible to severe growth reductions by Swiss
needle cast, that warmer spring and summer
temperatures are associated with Swiss needle
cast impacts, and that the disease appears to be
increas- ing in severity” (p. 1673).

Fomes pini
Older stands of Douglas-fir are subject to the
foliar and root diseases discussed earlier, but, as
trees age the incidence of bole rotting fungi
increases in oc- currence and virulence, as
discussed by Boyce and Wagg (1953). Much of
the following is based on their report. They noted
that “decay caused by Fomes pini has long been
referred to as conk rot, because this fungus is
unique in producing numerous conks, or fruiting
bodies, which follow closely the progress of rot in
Most of the other decay-causing fungi appear later in
growth trees examined in their study had 80% the life of a stand. As a stand approaches stagnation
to 90% decay by F. pini, making them and then declines, brown cubical rots, particularly
brown trunk rot caused by the quinine fungus, Fomes
effectively mas- sive columns of decay laricis, become increasingly important. Fomes laricis
surrounded by relatively thin layers of attacks not only sound trees but those already infected
with Fomes pini. In the latter instance, brown trunk
sapwood and bark. Such structures are subject
rot is able to overrun conk rot. Fomes laricis
to windthrow and bark beetle damage (Isaac commonly follows the major
1956, Stathers et al. 1994). Fomes pini, which attack of Fomes pini, and is the most important
causes a red ring rot, is responsible for 81% of fungus in the final break up of old-growth Douglas-
fir stands. Polyporus schweinitzii, depending on
total board foot volume of decay in western wounds for entrance, may appear at any time during
Oregon and Washington, and can attack trees stand devel- opment. Its incidence is proportional to
as young as 27 years (Boyce 1932, p. 33). the amount of
basal wounding. (Boyce and Wagg 1953, p. 10)
Boyce and Wagg (1953) noted that “Fomes
pini causes more decay on warmer sites” (p. The incidence of Fomes pini was higher on
70); accordingly, global warming may well good sites than poor, possibly because good sites
cause an increase in this fungus. According to have fewer trees with larger branches which offer
Boyce and Wagg (1953), infec- tion sites of heartwood when the branches
die. Pure stands of Douglas-fir had a higher
Several fungi are responsible for decay or rot in
Douglas- fir. A previous investigation showed
percentage of Douglas-fir infected with Fomes
that 80.8 per cent of the total board foot volume pini than did mixed stands. Boyce and Wagg
of decay found in Douglas- fir on plots in (1953, p. 89) observed that “Fomes pini is
western Oregon and Washington was red ring rot
or conk rot caused by Fomes pini . . . Nearly all somewhat pathogenic, commonly encroaching on
the remainder was divided among red-brown butt the sapwood, resulting either in the death of the
rot, caused by Polyporus schweinitzii; brown trunk tree directly or, as seems more likely, reducing its
rot, caused by Fomes laricis; and yellow-brown
top rot, caused by Fomes roseus—all three being vigor so that it succumbs to competition. The most
brown cubical rots. (Boyce and Wagg 1953, p. 7). rapidly growing trees are infected first, their
growth is reduced, and finally they drop out of
the stand. Meanwhile, new infections are was high, whereas salal, twinflower and rhododendron
occurring in the remaining trees, and the process indicated a lower incidence of decay.
Pure stands contained greater volumes of conk rot
is repeated” (p. 89). Their conclusions were as than did mixed stands; stands of poor quality were
follows: more defective than those of good quality; and stands
that had been damaged by fire had more conk rot than
No relationship was found between the number of those that had not been burned. (Boyce and Wagg
trees with conk rot in a stand and such factors as 1953, p. 90)
elevations ranging from 1,500 to 4,500 feet, curvature
of slope, texture of the soil, and acidity of the soil. A
greater amount of conk rot occurred in trees on steep
Conclusion
slopes, on southerly aspects, on upper slopes or Although this chapter has provided an
hogbacks, on soil with excessive drainage, and on
shallow soil.
overview, additional ontogenetic work is
In nearly all instances higher temperature is needed to continue to increase our understanding
associ- ated with the foregoing factors, so possibly of the morphological, anatomical, and
temperature is the controlling factor. The incidence of
conk rot was lower in stands that were on moderate
physiological changes that occur as Douglas-fir
slopes, on north- erly aspects, on lower slopes and develops from a seedling to a mature tree.
benches, on soils with good to restricted drainage, and Finally, we must continue to expand our un-
on deep soils. Soils with a high percentage of total
nitrogen produced more decayed trees than those with derstanding of the environmental and other
a low percentage. Where vine maple, vanillaleaf, factors that can affect this important genus at its
oxalis, or rose predominated in the secondary different life stages.
vegetation, the incidence of conk rot
References
Aagaard, JE, SS Vollmer, FC Sorensen, and SH Strauss. eco-
1995. Mitochondrial DNA products among RAPD
profiles are frequent and strongly differentiated between
races of Douglas-fir. Molecular Ecology 4:441–447.
Abbot, HG. 1973. Direct seeding in the United States, pp. 1–
9 in Direct Seeding Symposium, September 4–13, 1973,
Timmins, Ontario, ed. JH Crayford. Canadian Forestry
Service, Ministry of Natural Resources Publication No.
1339, Ottawa, ON.
Abele, NI. 1909. Das waldbauliche Verhalten der Dougla-
sien. Naturwissenschaftliche Zeitschrift fur Forst- und
Land- wirtschaft 7(9):477–479.
Ackerman, RF, and LL Farrar. 1965. The Effect of Light and
Temperature on the Germination of Jack Pine and
Lodgepole Pine Seeds. Technical Report 5, Faculty of
Forestry, Uni- versity of Toronto, ON.
Adams, DL. 1981. The northern Rocky Mountain region, pp.
341–389 in Regional silviculture of the United States, John
W. Barrett, ed. John Wiley, New York.
Adams, GW, and MS Greenwood. 1992. Optimization of
environmental regimes for flowering in an indoor
breed- ing hall for black spruce, white spruce, and
jack pine, pp. 219–227 in Proceedings AFOCEL/IUFRO
Conference, Bordeaux, France. Vol. I. AFOCEL 77370,
Nancy, France.
Adams, WT. 1981. Population genetics and gene
conservation in Pacific Northwest conifers, pp. 401–415
in Evolution Today: Proceedings of the Second
International Congress of Systematic and Evolutionary
Biology, July 17–24, 1980, ed. GGE Scudder and JL
Reveal. University of British Columbia, Vancouver, BC.
Adams, WT. 1982. Clonal variation in pollen-related charac-
teristics of Douglas-fir. Canadian Journal of Forest
Research 12:403–408.
Adams, WT, and JC Bastien. 1994. Genetics of second flush-
ing in a French plantation of coastal Douglas-fir. Silvae
Genetica 43:345–352.
Adams, T, and RK Campbell. 1982. Genetic adaptation and
seed source specificity, pp. 78–85 in Reforestation of
Skel- eton Soils symposium, ed. OT Helgerson and SD
Hobbs, Forest Research Laboratory, Oregon State
University, Corvallis, OR.
Adams, WT, and DG Joyce. 1990. Comparison of selection
methods for improving volume growth in young coastal
Douglas-fir. Silvae Genetica 39 (5/6):219–226.
Adams, WT, DS Birkes, and VJ Erickson. 1992. Using
genetic markers to measure gene flow and pollen
dispersal in forest tree seed orchards, pp. 37–61 in
Ecology and Evo- lution of Plant Reproduction, ed R
Wyatt, Chapman and Hall, New York.
Addicott, FT, JL Lyon, K Ohkuma, WE Thiessen, HR Carns,
OE Smith, JW Cornforth, BV Milborrow, G Ryback,
and PF Wareing. 1968. Abscisic Acid: A New Name for
Abscisin II (Dormin). Science 159(3822):1493.
Adolph, L. 1936. Der Grossanbau der Douglasie im
Gadower
Forst. Forstarchiv 12:183–191.
Adolph, L. 1949. Die Douglasie. Forstwirtschaft–
Holzwirtschaft
3(15):242–248.
Aft, H. 1961. Chemistry of dehydroquercitin I. Acetate
deriva- tives. Journal of Organic Chemistry 26:1958–
1963. 285
Agee, JK. 1994. Fire and weather disturbances in terrestrial
systems of the eastern Cascades. USDA Forest Service,
Gen- eral Technical Report GTR-PNW-320, Pacific
Northwest Research Station, Portland, OR.
Agee, JK, and MH Huff. 1987. Fuel succession in a western
hemlock/Douglas-fir forest. Canadian Journal of Forest
Research 17:697–704.
Agee, JK, M Finney, and R De Gouvenain. 1990. Forest his-
tory of Desolation Peak, Washington. Canadian Journal
of Forest Research 201:350–356.
Aiden, D. 1985. Biology and management of white spruce
seed crops for reforestation in subarctic Taiga forests.
Bulletin 69, Agriculture and Forestry Experiment
Station, School of Agriculture and Land Resources
Management, University of Alaska-Fairbanks.
Aitken, SN, and WT Adams. 1996. Genetics of fall and
winter cold hardiness of coastal Douglas-fir in Oregon.
Canadian Journal of Forest Research 26(10):1828–
1837.
Aitken, SN, and WT Adams. 1997. Spring cold hardiness
under strong genetic control in Oregon populations of
Pseudotsuga menziesii var. menziesii. Canadian Journal
of Forest Research 27(11):1773–1780.
Aitken, SN, WT Adams, N Schermann, and LH Fuchigami.
1996. Family variation for fall cold hardiness in two
Wash- ington populations of coastal Douglas-fir
(Pseudotsuga menziesii var. menziesii (Mirb.) Franco.
Forest Ecology and Management 80:187–195.
Akimockin, NG. 1964. Pseudotsuga taxifolia on
unpromising sites at the Forest-Steppe Experiment
Station. Lesnoi Zhurnal 7(2):35–38.
Alden, JN. 1971. Freezing resistance of tissues in the twig of
Douglas-fir (Pseudotsuga menziesii (Mirb.) Franco.
Doctoral dissertation, Oregon State University,
Corvallis. 149 p.
Alexander, RR. 1974. Silviculture of Subalpine Forests in
the Central and Southern Rocky Mountains: The Status
of Our Knowledge. USDA Forest Service, Research Paper
RM-121, Rocky Mountain Forest and Range Experiment
Station, Fort Collins, CO. 88 p.
Alexander, RR. 1985. Major Habitat Types, Community
Types, and Plant Communities in the Rocky Mountains.
USDA Forest Service, General Technical Report RM-
123. Rocky Mountain Forest and Range Experiment
Station, Fort Collins, CO. 105 p.
Alexander, RR. 1988. Forest vegetation on National Forests in
the Rocky Mountain and Intermountain regions: habitat
types and community types. USDA Forest Service, General
Techni- cal Report RM-162. Rocky Mountain Forest and
Range Experiment Station, Fort Collins, CO. 47 p.
Allan, CPS. 1978. Position statement-Westland
Conservancy, pp. 336–340 in A Review of Douglas Fir
in New Zealand, ed. RN James and EH Bunn. FRI
Symposium No. 15, Forest Service, Forest Research
Institute.
Allegri, E. 1962. La introduzione e la sperimentazione in
Italia di specie forestali esotiche a rapido accrescimento.
Monti e Boschi 13(11/12):507–519.
Allen, GS. 1941a. A basis of forecasting seed crops of some
coniferous trees. Journal of Forestry 39:1014–1016.
Allen, GS. 1941b. A standard germination test for Douglas-fir
seed. Forestry Chronicle 17:75–79.
Allen, GS. 1941c. Light and temperature as factors in the
ger-
mination of the seed of Douglas-fir (Pseudotsuga taxifolia
286 Douglas-fir: The Genus Pseudotsuga
(Lamb.) Britt). Forestry Chronicle 17:99–109. survival of exotic species at Cloquet, Minnesota. Minnesota
Allen, GS. 1943. The embryogeny of Pseudotsuga Forestry Research Notes No. 237. College of Forestry,
taxifolia University of Minnesota, St. Paul.
(Lamb.) Britt. American Journal of Botany 30(8):655– Alosi, MC, and DB Neale. 1992. Light- and phytochrome-
661. Allen, GS. 1946a. Embryogeny and development of mediated gene expression in Douglas-fir seedlings. Phys-
the api-
cal meristems of Pseudotsuga. I Fertilization and early
embryogeny. American Journal of Botany 33(8):666–
677. Allen, GS. 1946b. The origin of the microsporangium of
Pseu-
dotsuga. Bulletin of the Torrey Botanical Club 73(6):547–556.
Allen, GS. 1947a. Embryogeny and the development of the
apical meristems of Pseudotsuga II. Late embryogeny.
American Journal of Botany 34:73–80.
Allen, GS. 1947b. Embryogeny and the development of the
apical meristems of Pseudotsuga III. Development of the
apical meristems. American Journal of Botany 34:204–
211.
Allen, GS. 1947c. Mold free germination of coniferous
seeds.
Journal of Forestry 45:51.
Allen, GS. 1957a. Better handling of a scarce commodity.
British Columbia Lumberman 41(7):32–36.
Allen, GS. 1957b. Storage behavior of conifer seeds in sealed
containers held at 0°F, 32°F, and room temperature.
Journal of Forestry 55:278–280.
Allen, GS. 1958a. Factors affecting the viability and
germina- tion behavior of coniferous seed. Part III.
Commercial processing and treatments similar to
processing; Pseu- dotsuga menziesii (Mirb.) Franco, and
other species. Forestry Chronicle 34(3):266–298.
Allen, GS. 1958b. Factors affecting the viability and
germina- tion behavior of coniferous seed. Part II. Cone
and seed maturity, Pseudotsuga menziesii (Mirb) Franco.
Forestry Chronicle 35(3):275–282.
Allen, GS. 1960a. Factors affecting the viability and
germina- tion behavior of coniferous seed. Part IV.
Stratification period and incubation temperature,
Pseudotsuga menziessi (Mirb) Franco. Forestry
Chronicle 36(1):18–29.
Allan, GS. 1960b. A Method of Distinguishing Coastal and
Interior Douglas-fir Seed. University of British Columbia,
Faculty of Forestry Research Note No. 28.
Allen, GS. Testing Douglas-fir seed for provenance.
Proceedings of the International Seed Testing Association
26(3):388–403.
Allen, GS. 1962a. Factors affecting the viability and
germina- tion behavior of coniferous seed. V. Seed
moisture content during stratification and secondary
storage. Pseudotsuga menziesii (Mirb) Franco. Forestry
Chronicle 38(3):303–308.
Allen, GS. 1962b. Factors affecting the viability and germi-
nation behavior of coniferous seed. VI. Stratification
and subsequent treatment, Pseudotsuga menziesii (Mirb.)
Franco. Forestry Chronicle 38(4):485–496.
Allen, GS. 1962c. The deterioration of Douglas-fir seed
under
various conditions. Forestry Chronicle 38(1):145–147.
Allen, GS. 1963. Origin and development of the ovule in
Douglas-fir. Forest Science 9(4):386–393.
Allen, GS. 1967. Stratifications of tree seed. The International
Plant Propagation Society Combined Proceedings 17:99–106.
Allen, GS, and W Bientjes. 1954. Studies on coniferous seed
at the University of British Columbia. Forestry
Chronicle 30(2):183–196.
Allen, GS, and JN Owens. 1972. The Life History of
Douglas- fir. Canadian Forestry Service, Environment
Canada. Ottawa, ON. 139 p.
Allen, MF. 1991. The Ecology of Mycorrhizae. Cambridge
Uni- versity Press, Cambridge.
Alm, AA, RA Jensen, and BA Brown. 1972. Growth and
iologia Plantarum 86:71–76. fremdländischen Holzarten. Allgemeine Forst und
Alosi, MC, DB Neale, and CS Kinlaw. 1990. Expression Jagdzeitung 80:371–374.
of cab genes in Douglas-fir is not strongly regulated Anonymous. 1954. Rules for testing seeds. Proceedings As-
by light. Plant Physiology 93:829–832. sociation of Official Seed analysts 44:31–78.
Alten, P von. 1898. Die Einbürgerung fremder Baumarten Anonymous. 1965. New South Wales steps up Douglas-fir
in Deutschland. Vortrag Wiesbaden. Bechtold & Cie.
Alvarez, AD. 1994. Analisis historico–ecologico de los
bosques de Pseudotsuga en Mexico. Folieto Tecnico
No. 24. Secretaria de Agricultura y Recursos
Hidraulicos, Instituto Nacional de Investigaciones
Forestales y Agropecurias, Centro de Investigacion
Regional del Golfo Centro Campo Experi- mental el
Palmar, Ver. Mexico.
Alyarez, IF, and RG Linderman. 1983. Effects of
ethylene and fungicide dips during cold storage on
root regeneration and survival of western conifers
and their mycorrhizd fungi. Canadian Journal of
Forest Research 13(5):962–971.
Amaranthus, MP. 1994. Mycorrhizas, forest
disturbance and regeneration in the Pacific
northwestern United States, pp. 202–207 in
Mycorrhizas in Ecosystems, ed. DJ Read, DH
Lewis, AH Fitter, and IJ Alexander. CAB
International, Singapore.
Amaranthus, MP, and DA Perry. 1987. Effect of soil
transfer on ectomycorrhizal formation and the
survival and growth of conifer seedlings on old,
non-reforested clear-cuts. Canadian Journal of
Forest Research 17:944–950.
Amaranthus, MP, and DA Perry. 1989. Interaction
effects of vegetation type and Pacific madrone soil
inocula on sur- vival, growth and mycorrhizal
formation on Douglas-fir. Canadian Journal of
Forest Research 19:550-556.
Amaranthus, MP, D Page-Dumroese, A Harvey, E
Cazares, and LF Bedman. 1996. Soil Compaction and
Organic Matter Affect Conifer Seedling Non-
Mycorrhizal and Ectomycorrhizal RootTip Abundance
and Diversity. USDA Forest Service Pacific
Northwest Research Station. Research Paper PNW-
RP-494. Portland, OR.
Anderson, ML. 1961. The Selection of Tree Species, an
Ecologi- cal Basis of Site Classification for
Conditions Found in Great Britain and Ireland. 2nd
ed. Oliver & Boyd, Edinburgh and London.
Anderson, ML. 1967. A History of Scottish Forestry.
Vol. 2. Thomas Nelson & Sons, London and
Edinburgh, Charles J Taylor.
Anekonda, TS, and WT Adams. 2000. Genetics of dark
res- piration and its relationship with drought
hardiness in coastal Douglas-fir. Thermochimica
aeta 349:69–77.
Anekonda, TS, WT Adams, and SN Aitken. 1998.
Influence of second flushing on genetic assessment
of cold hardiness in coastal Douglas-fir (Pseudotsuga
menziesii (Mirb.) Franco). Forest Ecology and
Management 111:119–126.
Anekonda, TS, WT Adams, SN Aitken. 2000. Cold
hardi- ness testing for Doulgas-fir tree improvement
programs: guidelines for a simple, robust, and
inexpensive screening methods. Western Journal of
Applied Forestry 15(3):129–136.
Anger, L. 1879. Abies Douglasii-Douglas Tanne.
Vereinsschrift für Forst-, Jagd- und Naturkunde
105:77–80.
Annila, E. 1984. Population fluctuation of some cone
and seed insects in Norway spruce, pp. 57–64 in
Proceedings of the IUFRO Cone and Seed Insects
Working Party Conference. Working Party S2.07–
01. July 31–August 6, 1983, Athens, GA, ed. HO
Yates, III. USDA Forest Service Southeastern
Forest Experiment Station, Asheville, NC. 214 p.
Anonymous. 1904. Versammlungen Norddeutscher
Forst- vereine im Jahre 1903. III. Ost– und
Westpreussischer Forstverein. Versuche mit
References 287
on Douglas-fir: pathogenicity and seed stratification method
planting. Australian Timber Journal 31(10):100. to decrease Fusarium contamination. New Forests 9:35–51.
Anonymous. 1966. Fire devastates large area in Smith River
country. Forest Log 36(2):1,4.
Anonymous. 1971. Ice from silver thaw causes breakage
in
forest stands. Forest Log 40(6):3.
Anonymous. 1985. Hawaii celebrates 50th anniversary
Doug- las-fir planting. Western Forester 30(4):13.
Anonymous. 1990. Forest Nutrition Research Project, Proposal
for five-year extension of research. College of Forest
Resources, University of Wash. 34 p.
Anonymous. 1994. A brief history of Douglas-fir in New
Zealand. New Zealand Forestry 39(1):28–30.
Apple, M, K Tiekotter, M Snow, J Young, A Soeldner, D
Phillips, D Tingey, and BJ Bond. 2002. Needle anatomy
changes with increasing tree age in Douglas-fir. Tree
Physiology 22(2–3):129–136.
Apsit, VJ, RC Nakamura, and NC Wheeler. 1989.
Differential male reproductive success in Douglas-fir.
Theoretical and Applied Genetics 77:681–684.
Arber, A. 1941. The interpretation of leaf and shoot in the
angiosperms. Biology Review 16:81–105.
Arenas, JC. 1962. Repoblaciones con Pseudotsuga douglasii
Carr., en la provincia de Gerona, pp. 540–546 in
Asamblea Tecnica Forestal 1962. Ministerio de Agricultura,
Direccion Gene- ral de Montes, Caza y Pesca fluvial,
Madrid.
Ares, A, TA Terry, RE Miller, HW Anderson, HW, and BL
Fleming. 2005. Ground based forest harvesting effects
on soil physical properties and Douglas-fir growth. Soil
Science Society of America 69:1822–1832.
Arno, SF. 1988. Fire ecology and its management
implication in ponderosa pine forests, pp. 133–140 in
Proceedings of a Symposium on ponderosa pine: The
species and its management, ed. D Baumgartner and J
Lotan. Cooperative Extension, Washington State
University, Pullman, WA.
Arno, SF, JH Scott, MG Hartwell. 1995. Age-class structure of
old-growth ponderosa pine/Douglas-fir stands and its relation
to fire history. USDA Forest Service Intermountain
Research Station, Research Paper INT-RP-481, Ogden, UT.
Arnold, CA. 1935. A Douglas-fir cone from the Miocene of
Southeastern Oregon. Journal of Washington Academy
of Science 25(8):378–380.
Arora, R, LJ Rowland, and K Tanino. 2003. Induction and
release of bud dormancy in woody perennials: A science
comes of age. HortScience 38(5):911–921.
Ascherson, P, and P Graeber. 1913. Synopsis der mitteleu-
ropäischen Flora. 2nd ed., vol. I. Wilhelm Engelmann,
Leipzig.
Atkin, RK, GE Barton, and DK Robinson. 1973. Effect of
root-growing temperature on growth substances in xy-
lem exudate of Zea mays. Journal of Experimental
Botany 24(2):475–487.
Aughanbaugh, J. 1960. Comparative Growth of Eight Conifers
in a Plantation in Mahoning County, Ohio. Special Circular
Ohio Agricultural Experiment Station No. 96.
Axelrod, DI. 1980. History of the Maritime Closed–cone Pines,
Alta and Baja California. University of California Publica-
tions Geological Sciences 120.
Axelrood, PE, D Fong, and R Radley. 1990. Incidence and
pathogenicity of seedborne Fusarium on Douglas-fir
(Pseu- dotsuga menziesii Mirb Franco). Phytopathology
80(10):1065.
Axelrood, PE, WK Chapman, KA Seifert, DB Trotter, and H
Shrimpton. 1998. Cylindrocarpon and Fusarium root
colo- nization of Douglas-fir seedlings from British
Columbia reforestation sites. Canadian Journal of
Forest Research 28:1198–1206.
Axelrood, PE, M Neumann, D Trotter, R Radley, G
Schrimpton, and JJ Dennis. 1995. Seedborne Fusarium
flowering and on nitro- gen and sugar composition of
Badoux, H. 1926. Observations sur le douglas vert en slash pine. Forest Science 14:172–180.
Suisse. Mitteilungen der Schweizerischen Barnes, V. 1985. Cone and seed handling and protection (on
Zentralanstalt für das forst- liche Versuchswesen numbered pages IV) D DeYoe, compiler. Cone and Seed
14(1):3–27. processing Workshop Proceedings (on numbered pages).
Bailey, JD, and CA Harrington. 2006. Temperature OSU Extension Service, OSU, Corvallis, OR.
regulation of bud-burst phenology within and Barnett, JP. 1980. Inhibition temperatures affect seed
among years in a young Douglas-fir (Pseudotsuga moisture uptake, germination and seedling development,
menziesii) plantation in western Washington, USA. pp. 41–45 in Southern Silviculture Research Conferences,
Tree Physiology 26:421–430. 6–7 November 1980, Atlanta, GA.
Baker, FS. 1929. Effect of excessively high temperature
on
coniferous reproduction. Journal of Forestry 27:949–
975.
Baker, FS. 1934. Theory and Practice of Silviculture.
McGraw- Hill, New York.
Baker, HG. 1972. Seed Weight in relation to
environmental conditions in California. Ecology
53(6):997–1010.
Baldwin, HI. 1942. Forest Tree Seed of the North
Temperate Re- gions: with Special Reference to North
America. Chronica Botanica, University of Michigan.
Baldwin, HI, and D Murphy. 1956. Rocky Mountain
Douglas- fir succeeds in New Hampshire. Fox
Forest Note 67, New Hampshire Forestry and
Recreation Commission, Caroline A. Fox
Research and Demonstration Forest, Hillsboro. 2
p.
Baldwin, JL. 1973. Climate of the United States. U.S.
Depart- ment of Commerce, National Oceanic and
Atmospheric Administration, Washington, DC. 47
p.
Ballian, D, T Mikić, and K Pintarić. 2002.
Provenijenični pokusi sa zelenom duglazijom
(Pseudotsuga menziesii (Mirb.) Franco) na
lokalitetu Blinje kod Kreševa. Provenance trials
with Douglas-fir at Blinje site near Kreševa, Works
of the Faculty Forestry, University Sarajevo. Nr
1:9-18.
Ballian, D, T Mikić, K Pintarić, and M Šćekić. 2003.
Analiza rasta zelene duglazije (Pseudotsuga
menziesii (Mirb.) Fran- co). IUFRO pokusu
“Gostovic” Zavidovici. With English summary.)
[The analysis of Douglas-fir (Pseudotsuga
menziesii (Mirb.) Franco grown within IUFRO
trials at “Gostovic” Zavidovici. Works of the
Faculty of Forestry, University of Sarajevo 1:55-
63.]
Bano, I. 1963. A duglasfenyö–allomanyok
erdömüvelese hazai tapasztalatok alapjan. (with
German summary: Der Waldbau der
Douglasienbestände auf Grund ein- heimischer
Erfahrungen). Magyar Tudomanyos Akademia
Agrartudomanyok Osztalyanak Közlemenyei 22(1/2):93–
104.
Barber, A. 1977. Nutrients in the soil and their flow to
plant roots, pp. 161–169 in The Belowground
Ecosystem: A Synthe- sis of Plant-associated
Processes, ed. JK Marshall. Colorado State University
Range Science Department, Science Series No. 26,
Fort Collins, CO.
Barber, A. 1986. Lillooet Lake Douglas-fir Plantation
Shading Trail Fall 1987 Analysis. BSF Essay, Faculty
of Forestry, University of British Columbia,
Vancouver, BC.
Barner, H. 1973. Procurement of Douglas-fir seed for
prov- enance research, pp. 82–87 in Proceedings
Meeting of IU- FRO Working Party on Douglas-fir
provenances S2.02–05, September 3–5, H.H.
Hattemer ed. Göttingen, Germany.
Barner, H, and H Christianson. 1962. The formation
of pol- len, the pollination mechanism and the
determination of the most favorable time for
controlled pollination in Pseudotsuga menziesii.
Silvae Genetica 11(4):89–102.
Barnes, RL, and GW Bengston. 1968. Effects of
fertilization, irrigation and cover cropping on
288 Douglas-fir: The Genus Pseudotsuga
Barnett, JP and BF McLemore. 1984. Germination speed as Bedö, A. 1878. A douglasfenyö, Abies Douglasii (Lindley).
a predictor of nursery seedling performance. Southern Erdészeti Lapok 17:643–656.
Journal of Applied Forestry 8(3):154–162. Begin, J. 1992. Productivité du douglas vert (Pseudotsuga
Baron, FJ. 1971. Ten years of forest seed crops in California. menziesii (Mirb.) Franco var. menziesii Franco) en
Journal of Forestry 69(4):490-492. relation avec des caractéristiques stationelles. (with
Barrett, SW, SF Arno, and JP Menakis. 1997. Fire episodes Engl. abstract: Yield performance of Douglas-fir
in the Inland Northwest (1540-1940) based on fire (Pseudotsuga menziesii (Mirb.) Franco var. menziesii
history data. USDA Forest Service General Technical Franco) in relation to site characteristics. Mitteilungen der
Report INT-GTR-370, Intermountain Research Station, Eidgenössischen Forschun- gsanstalt für Wald, Schnee
Ogden, UT. 17 p. und Landschaft 67(2):173–313.
Barthelmess, A. 1935. Über den Zusammenhang zwischen Behler, H. 1980. Über die Zukunft der Douglasie aus forst-
Blattstellung und Stelenbau unter besonderer wirtschaftlicher und holzwirtschaftlicher Sicht.
Berücksi- chtigung der Koniferen. Botanisches Archiv. Allgemeine Forstzeitschrift 35(9/10):222–224.
37:207–260. Belcher, EW. 1982. Storing stratified seeds for extended peri-
Barton, LV. 1954a. Effect of sub-freezing temperatures on ods. USDA Forest Service Tree Planters’ Notes 33(4):23–
viability of conifer seeds in storage. Contribution Boyce 25.
Thompsen Institution 18(1):21–24. Bell, HE, RF Stettler, and RW Stonecypher. 1979. Family
Barton, LV. 1954b. Storage and packeting of seeds of and fertilizer interaction in one year old Douglas-fir.
Douglas- fir and Western hemlock. Contribution Boyce Silvae genetica 28:1–5.
Thompson Institution 18:25–38. Bellmann, E, and Schönbach, H. 1964. Erfolgsaussichten
Baskin, CC, and JM Baskin. 1998. Seeds: Ecology, der Auslesezüchtung auf Frostresistenz bei der grünen
Douglasie (Pseudotsuga menziesii (Mirb.) Franco).
Biogeography and Evolution of Dormancy and Archiv für Forstwesen 13(3):307–331.
Germination. Academic Press. New York.
Belton, MC. 1991. Forestry: a sustainable land use for high
Bastien, JC, Y Birot, and R Lanares. 1980. IUFRO Douglas- country lands. New Zealand Forestry 36(1):19–22.
fir provenances experiment. Peyrat-le-Chateau (Haute-
Vienne). INRA, Station d’Amélioration des Arbres Belton, ME, and MR Davis. 1986. Growth decline and
forestiers. Mimeo. phospho- rous response by Douglas-fir one degraded
high-country yellow-brown earth. New Zealand Journal
Bastien, JC, B Roman-Amat, and P Ferrandes. 1988. Faut-il of Forestry Sci- ence 16(1):55–68.
s’interésser au douglas bleu? Revue forestière française
40(6):436–446. Bennett, WH, and AP Long. 1919. Report of the judges on
the plantations competition held in connection with the
Bastien, JC, Y Birot, C Christophe, R Lanares, and M Royal Agricultural Society’s show at Cardiff 1919.
Faucher. 1980. IUFRO Douglas-fir provenance–progeny test. Quarterly Journal of Forestry 13:222–253.
Cendrieux (Dordogne). INRA, Station d’Amélioration
des Arbres forestiers. Mimeo. Berennier, G. 1988. The control of floral evocation and mor-
phogenesis. Annual Review of Plant Physiology and
Bates, CG. 1924. Forest types in the central Rocky Plant Molecular Biology 39:175–219.
Mountains as affected by climate and soil. USDA Berg, FV. 1912. Einige Beobachtungen aus der Baumzucht.
Bulletin 1233, Wash- ington, DC. Mitteilungen der Deutschen Dendrologischen Gesellschaft
Bates, SL, JH Bordon, AR Kermode, and RG Bennett. 21:55–67.
2000a. Impact of Leptoglossus occidentalis (Hemiptera: Bergen, JY. 1904. Elements of Botany, pp.17–18. Ginn &
Coreidae) on Douglas-fir seed production. Journal of Co., rev. ed. Boston and New York.
Economic En- tomology 93(5):1444–1451.
Bergmann, F. 1973. Analysis of genetic variation in a
Bates, SL, CG Lait, JH Borden, and AR Kermode. 2001. Douglas- fir provenance by means of isozyme
Effect of feeding by the western conifer seed bug, polymorphisms, pp. 207–215 in Proceedings IUFRO
Leptoglossus occidentalis on the major storage reserves S2.02–05 Working Party on Douglas-fir provenances,
of developing seeds and on seedling vigor of Douglas- Sept. 3–5, 1973. Göttingen, Germany.
fir. Tree Physiol- ogy 211:481–487. Bernetti, G. 1987. I boschi della Toscana. I quanderni di Monti
Bates, SL, JH Borden, H Savoie, SE Blatt, CG Lait, AR e Boschi. Edagricole, Bologna.
Kermode, and RG Bennett. 2000b. Impact of feeding by Berney, JLA. 1972. Studies on the probable origin of some Eu-
Leptoglossus occidentalis (Hemiptera: Coreidae) on the ropean Douglas-fir (Pseudotsuga menziesii (Mirb.)
major storage reserves of mature Douglas-fir (Pinaceae) Franco) plantations. MF thesis, University of British
seeds. The Canadian Entomologist 132:91–102. Columbia, Vancouver, BC.
Bayerische Staatsministerium für Ernährung, Landwirtschaft Bernier, G. 1988. The control of floral evocation and mor-
und Forsten, Ministerialforstabteilung. 1970. phogenesis. Annual Review of Plant Physiology and
Zusammen- stellung der Douglasienvorkommen im Plant Molecular Biology 39, 175–219.
Staatswald nach Ober- forstdirektionen, Wuchsgebieten
und Altersklassen. Beilage 3 zur MFE vom 12.3.1970, Bernier G, JM Kinet, and RM Sachs. 1981. The physiology of
No. F 2306/70–OD 850 b. flowering, Vol II. The initiation of flowers. CRC Press Inc.,
Boca Raton, FL.
Beckwith, RC. 1978. Biology of the insect, pp. 25–29, in The
Bernier, PY, MS Lamhamedi, and DG Simpson. 1995.
Douglas-fir Tussock Moth: A Synthesis, ed. MH Shoot:root ratio is of limited use in evaluating the quality
Brookes, RW Stark, and RW Campbell. USDA Forest of container conifer stock. Tree Planters’ Notes 46:102–
Service, Science and Education Agency, Technical 105.
Bulletin 1585. Washington, DC. 331 p.
Bernier, G, A Havelange, C Houssa, A Petitjean, and P
Bedard, WD. 1938. An annotated list of the insect fauna of Lejeune. 1993. Physiological signals that induce
Douglas fir (Pseudotsuga mucronata Rafinesque) in the flowering. Plant Cell 5:1147–1155.
northern Rocky Mountain Region. The Canadian Ento-
mologist 70(9):188–197. Bevan, D. 1967. Forest entomology. Report on Forest Research,
London 1967:103–105.
Bedell, JP, M Chalot, A Garnier, and B Botton. 1999. Effects
of nitrogen source on growth and activity of nitrogen- Bevan, D. 1968. Forest entomology. Report on Forest Research,
assimilating enzymes in Douglas-fir seedlings. Tree London 1968:112–115.
Physi- ology 19:205–210. Bevan, D. 1969. Douglas-fir Seed Wasp Megastigmus spermot-
mophus. Report on Forest Research London 1969:111–
113.
References 289
ed. GA Lang.
Bever, DN. 1954. Evaluation of Factors Affecting Natural
Produc- tion of Forest Trees in Central Western Oregon.
Research Bulletin 3. Oregon State Board of Forestry.
Bevins, CD. 1980. Estimating survival and salvage potential
of fire-scarred Douglas-fir. USDA. Forest Service, Inter-
mountain Forest & Range Experiment Station, Research
Note INT-287, Ogden, UT. 8 p.
Bewley, JD. 1997. Seed Germination and Dormancy. The Plant
Cell 9(7):1055–1066.
Bewley JD, and M Black. 1978. Physiology and biochemistry of
seeds in relation to germination. Vol. 1. Development,
germi- nation and growth. Springer-Verlag, Berlin.
Bewley JD, and M Black. 1982. Physiology and biochemistry of
seeds in relation to germination. Vol. 2. Springer-Verlag,
New York.
Bewley, JD, and M Black. 1994. Seeds: Physiology of
Development and Germination, 2nd ed. Plenum Press, New
York. 445 p.
Binder, WD, and P Fielder. 1996. Chlorophyll fluorescence
as an indicator of frost hardiness in white spruce (Picea
glauca [Moench.] Voss) seedlings from different
latitudes. New Forests 11:233–253.
Binder, WD, P Fielder, GH Mohammed, and SJL
L'Hirondelle. 1997. Applications of chlorophyll
fluorescence for stock quality assessment with different
types of fluorometers. New Forests 13:63–89.
Black, BA, DC Shaw, and JK Stone. 2010. Impacts of Swiss
needle cast on overstory Douglas-fir forests of the west-
ern Oregon Coast Range. Forest Ecology and
Management 259:1673–1680.
Black, M. 1998. Knocking ’em dead. Biologist 45:99.
Black, M and JD Bewley, eds. 2000. Seed Technology and Its
Bio- logical Basis. Sheffield Academic Press Ltd.,
Sheffield, UK.
Black, M. KJ Bradford, and J Vázquez-Ramos, eds. 2000. Seed
Biology Advances and Applications. Proceedings of the
Sixth International Workshop on Seeds, Mérida, México,
1999. CABI International, New York.
Bialobok, S. 1959. Ausländer–Holzarten auf Versuchsflächen
in Polen. Archiv für Forstwesen 8(10):866–884.
Bialobok, S, and H Chylarecki. 1965. Badania nad upra a
drzew obcego pochodzenia w Polsce w warunkach
srodowiska lesnego (with Engl. summary: Cultivation of
exotic tree species in forest conditions in Poland).
Arboretum Kornickie Rocznik 10:211–276.
Bialobok, S, and L Mejnartowicz. 1970. Provenance differen-
tiation among Douglas-fir seedlings. Arboretum
Kornikie Rocznik 15:197–219.
Bianco J, G Garello, and MT Le Page-Degivry. 1997. De
novo ABA synthesis and expression of seed dormancy in
a gymnosperm: Pseudotsuga menziesii. Journal of Plant
Growth Regulation 21:115–119.
Bieler, R. 1935. Dotychczasasowe wyniki aklimatyzacji
drzew zagranicznych w Wielkopolsce i na kresach (Les
résultats actuels de l’acclimatation des arbres étrangers
en Haute– Pologne et dans la région septentrionale de
Pologne). pp. 148–153 in Proceedings of the First
Polish Symposium of Forestry Sciences, Poznan 1935.
University of Poznan Printing Office.
Bierhorst, DW. 1971. Morphology of Vascular Plants.
Macmillan
Co., New York.
Bigg, WL, and JW Schalau. 1990. Mineral nutrition and the
target seedling, pp. 139–160 in Target Seedling
Symposium: Proceedings, Combined Meeting of the Western
Forest Nursery Associations, August 13-17, 1990.
Roseburg, Oregon, ed. R Rose, SJ Campbell, and TD
Landis, USDA Forest Service, General Technical Report
RM-200, Roseburg, OR. 286 p.
Bigras, FJ. 1996. Conifer bud dormancy and stress
resistance: A forestry perspective, pp. 171–192 in Plant
Dormancy. Physiology, biochemistry and molecular biology.
Blake, J. 1987. The impact of a seedling diameter and root
CAB International Perspective, Oxon, UK. mass on performance of Douglas-fir and associated
conifers and their economic significance in
Binder, WD, and DJ Ballantyne. 1975. The respiration regeneration manage- ment. Unpublished.
and fertility of Pseudotsuga menziesii (Douglas fir)
pollen. Ca- nadian Journal of Botany 53(9):819–823. Blake, J, and WK Ferrell. 1977. The association between
doi 10.1139/b75-099. soil and xylem water potential, leaf resistance and
abscisic acid
Binder, WD, and P Fielder. 1990. Temperature and time-
related variation of root growth in some conifer tree
species. Canadian Journal of Forest Research
20:1192–1198.
Binder, WD, P Fielder, R Scagel, and GJ Krumlik.
1990. Tem- perature and time-related variation of
root growth in some conifer tree species. Canadian
Journal of Forest Research 20:1192–1199.
Binder, DB, GM Mitchell, and DJ Dallantyne. 1974.
Pollen vi- ability testing, storage and related
physiology. Canadian Forestry Service, Pacific
Forest Research Center Informa- tion Report No.
BC-X-105, Victoria, BC.
Binkley, D. 1981. Tree nutrition and carbohydrate
allocation to roots and shoots. Doctoral
dissertation, Oregon State University, Corvallis.
Birchler, TM, R Rose, and DL Haase. 2001. Fall
fertiliza- tion with N and K: Effects on Douglas-fir
seedling qual- ity and performance. Western
Journal of Applied Forestry 16(2):71–79.
Bird, CD, and WS Hong. 1969. Hepaticae of
Southwestern Alberta. Canadian Journal of Botany
47(11):1727–1746.
Birot, Y. 1972. Variabilité intraspécifique du poids de la
graine chez le douglas (Pseudotsuga menziesii Mirb.
Franco) Silvae Genetica 21(6):230–243.
Birot, Y, and G Burzynski. 1981. Analyse comparée
d’un test de provenances de douglas installé en
France et en Pologne. Revue forestière française
33(2):116–126.
Birot, Y, and P Ferrandes. 1972. Quelques aspects de la
varia- bilité du douglas (Pseudotsuga menziesii
Mirb.) introduit en zone mediterranéenne
subhumide. Ann. Sci. forest 29(3):335–351.
Birot, Y, and P Ferrandes. 1980. Douglas-fir breeding
pro- gramme in the French Mediterranean area.
INRA, Station d’Amélioration des Arbres
forestiers. Mimeo.
Birot, Y, and R Lanares. 1980. Douglas-fir
provenance expe- riment 1965. Peyrat-le-Chateau
(Haute-Vienne). INRA, Station d’Amélioration
des Arbres forestiers. Mimeo.
Biryukov, VT. 1971. (Introduction of Pseudotsuga
menzie- sii in the central regions of the RSFSR).
Lesnoi Zhurnal 14(5):8–10.
Bjorkman, E. 1970. Forest Tree Mycorrhizae-the
conditions for its formation and the significance
for tree growth and forestation. Plant and Soil
32(3) 589–610.
Black, HC, and WH Lawrence. 1992. Animal damage
man- agement in Pacific Northwest forests: 1901–
1990 in Silvi- cultural approaches to animal damage
management in Pacific Northwest Forests, ed. HC
Black. USDA Forest Service, Pacific Northwest
Research Station.
Black, HC, EJ Dimock II, J Evans, and JA Rochelle. 1979.
Animal damage to coniferous plantations in Oregon
and Washington. Oregon State University, School of
Forestry, Research Bulletin 25. Corvallis. 45 p.
Blackmore, S, and RB Knox. 1990. Microsporogenesis:
ther- mal program of development, pp. 1–10 in
Microspores: Evolution and Ontogeny. Blackmore,
S, and RB Knox eds. Academic Press, London and
Sydney.
Blackmore, S, CA McConchie, and RB Knox. 1988.
Phylogenetic analysis of the male ontogenetic
programme in aquatic and terrestrial
monocotyledons. Cladistics 3, 333–47.
290 Douglas-fir: The Genus Pseudotsuga
content in droughted seedlings of Douglas-fir Station, Fort Collins, CO.
(Pseudotsuga
menziesii). Physiologia Plantarum 39:106–109.
Blake, JI, and RH Linderman. 1992. A note on root devel-
opment, bud activity, and survival of Douglas-fir and
survival of western hemlock and noble-fir seedlings
fol- lowing exposure to ethylene during cold storage.
Canadian Journal of Forest Research 22:1995–2000.
Blake, J, SR Webster, and SP Gessel. 1988. Soil sulfate-
sulfur and growth responses of nitrogen-fertilized.
Douglas-fir to sulfur. Soil Science Society of America
Journal 52:1141–1147.
Blake, J, JB Zaerr, and S Hee. 1979. Controlled moisture
stress to improve cold hardiness and morphology of
Douglas-fir seedlings. Forest Science 25(4):576–582.
Blake, TJ, E Bevilacqua, GA Hunt, and SR Abrams. 1990.
Effects of abscisic acid and acetylenic alcohol and
dormancy, root development and transpiration in three
conifer species. Physiologia Plantarum 80:371–378.
Blanco, CE. 1941. Los pinos de Mexico. Mexico Forestal
19(1/2):7–10.
Blatt, SE, and JH Borden. 1996. Distribution and impact of
Leptoglossus occidentalis Heidemann (Hemiptera: Core-
idae) in seed orchards in British Columbia.The
Canadian Entomologist 128:1065–1076.
Blatt, SE, and JH Borden. 1998. Interactions between the
Douglas-fir seed chalcid, Megastigmus spermotrophus (Hy-
menoptera: Torymidae), and the western conifer seed
bug, Leptoglossus occidentalis (Hemiptera: Coreidae)
The Canadian Entomologist 130:775–782.
Bledsoe, CS, R Tennyson, and W Lopushinsky. 1982.
Survival and growth of outplanted Douglas-fir seedlings
inocu- lated with mycorrhizal fungi. Canadian Journal
of Forest Research 12:720–722.
Blenis, PV. 1995. Pathological problems of young trees
mod- ule 1, Forest ecology principles, environmental
training center land and services. Canadian
Department of En- vironmental Protection, Hinton,
AB.
Bloch, R. 1943. Polarity in plants. Botanical Reviews 9:261–
310. Bloomberg, WJ. 1966. The occurrence of endophytic
fungi
in Douglas-fir seedlings and seed. Canadian Journal of
Botany 44(4):413–420.
Bloomberg, WJ. 1969. Disease of Douglas-fir seeds during
cone storage. Forest Science 15(2):176–181.
Bloomberg, WJ. 1970. Seed disease and Douglas-fir during
cone storage. Bi-Monthly Forest Research Notes
Department of Fisheries and Forestry 26(2):15–16.
Bloomberg, WJ. 1973. Fusarium root rot of Douglas-fir
seed- lings. Phytopathology 63:337–341.
Bloomberg, WJ, and W Lock. 1974. Importance of
treatment timing in the control of Fusarium root rot of
Douglas-fir seedlings. Phytopathology 64:1153–1154.
Bloomberg, WJ, and WR Orchard. 1969. Chemical control
of root disease of Douglas-fir seedlings in relation to
fungus and nematode populations. Annals of Applied
Biology 64:239–244.
Bloomberg, WJ, and G Reynolds. 1982. Factors affecting
transfer and spread of Phellinus weirii mycelium in roots
of second growth Douglas-fir. Canadian Journal of
Forest Research 12(2):424–427.
Bloomberg, WJ, and GW Wallis. 1979. Comparison of
indica- tor variables for estimating growth reduction
associated with Phellinus weirii root rot in Douglas-fir
plantations. Canadian Journal of Forest Research
9(1):76–81.
Boden, F. 1902. Kritische Betrachtung ausländischer
Holzarten. Ein Beitrag zur Ausländerfrage.
Forstwissenschaftliche, Centralblatt 24:445–473, 542–
566, 601–616.
Boe, KN. 1953. Western Larch and Douglas-fir seed
dispersal into clear cuttings. Research Note 129.
Northern Rocky Mountain Forest and Range Experiment
Böhm, B. 1922. Ergebnisse des Anbaus ausländischer in Bezug auf ihren forstlichen Anbau in Deutschland. Julius
Hol- zarten. Ein Beitrag zur Ausländerfrage. Springer, Berlin.
Mitteilungen der Deutschen Dendrologischen Booth, J. 1880. Feststellung der Anbauwürdigkeit ausländischer
Gesellschaft 32:194–210. Waldbaeume. Referat auf Veranlassung der Königlich
Bohart, GE, and TW Koerber. 1972. Insects and seed Preus- sischen Hauptstation für forstliches Versuchswesen
produc- tion, pp. 1–53 in Seed Biology: Insects, and bearbeitet für die Versammlung des Vereins Deutscher
Seed Collection, Storage, Testing, and Certification, forstlicher Ver-
Volume 3, ed. TT Ko- zlowski, Academic Press,
NY.
Boiselle, R. 1953. Die Snoqualmie Douglasie–die Douglasie der
Zukunft. Allgemeine Forst– und Jagdzeitung 125(2):61–69.
Bold, HC, CJ Alexopoulos, and T Delevoryas. 1980.
Morphology of Plants and Fungi. Harper & Row,
New York.
Bolff, M and O Jenj. 1941. Dor Megastigmus Frage
Zeitschrift für Forst und Jagdwesen, pp. 113–137.
Bond, BJ. 2000. Age-related change in photosynthesis of woody
plants. Trends in Plant Science 5(8) 349–353.
Bond, BJ, and MH Ryan. 2000. Comment on:
Hydraulic limi- tation of tree height. A Critique
by Becker, Meinzer, and Wutlscheiger.
Functional Ecology 14:135–140.
Bond, BJ, NM Czarnomski, C Cooper, ME Day, and
MS Green- wood. 2007. Developmental decline in
height growth in Douglas-fir. Tree Physiology
27:441–453.
Bond, BJ, BT Farnsworth, RA Coulombe, and WE
Winner. 1999. Foliage physiology and
biochemistry in response to light gradients in
conifers with varying shade tolerance. Oecologia
120:183–192.
Bonnet-Masimbert, M. 1979. Flowering on Lammas
shoots of Douglas-fir, pp. 51–56 in Proceedings
Flowering and Seed Symposium, ed. F Bonner.
Mississippi State University Starksville.
Bonnet-Masimbert, M. 1982. Effect of growth
regulators, girdling, and mulching on flowering of
young European and Japanese larches under field
conditions. Canadian Journal of Forest Research
12:270–279.
Bonnet-Masimbert, M. 1982. Influence de l’ état
d’actavité des racines sur la floraison induite par des
gibberellines 4 et 7 chez Pseudotsuga menziesii
(Mirb.) Franco [Effect of root activity on flowering
stimulated by gibberellins 4 and 7 in Pseudotsuga
menziesii.] Silvae Genetica 31(5/6):178–183.
Bonnet-Masimbert, M. 1987. Floral induction in
conifers: a review of available techniques. Forest
Ecology and Man- agement 19:135–146.
Bonnet-Masimbert M, and P Doumas. 1992. Physiological
ap- proach to flowering induction in conifers, pp. 181–
188 in Proceedings AFOCEL/IUFRO Symposium on
Mass-production of genetically improved fast growing
forest tree species, Bor- deaux, France, September 14-
18, 1992, Vol. I.
Bonnet-Masimbert M, and R Lanares. 1978. Induction
florale surpousses d’aout chez le Douglas
(Pseudotsuga menziesii). Canadian Journal of
Forest Research 8:247–252.
Bonnet-Masimbert M, and JE Webber. 1995. From
flower induction to seed production in forest tree
orchards. Tree Physiology 15:419–426.
Bonnet-Masimbert M, and JB Zaerr. 1987. The role of
plant growth regulators in promotion of flowering,
pp. 13–35 in Hormonal Control of Tree Growth,
Kossuth, SV, and SD Ross eds. Martinus Nijhoff
Publishers, Boston, MA.
Booth, JR. 1841. Verzeichnis der bei der Versammlung
der deutschen Land– und Forstwirthe zu Doberan
im Sep- tember 1841 ausgestellten Nadelholzarten,
pp. 44–55 in Protokolle der Sitzungen der forstlichen
Sektion zu Doberan. Neue Jahrbücher der
Forstkunde, Heft 23. Darmstadt.
Booth, J. 1877. Die Douglas–Fichte und einige andere
Nadelhölzer namentlich aus dem nordwestlichen Amerika
References 291
Bracher, GA, PA Murtha. 1993. Relationship of the ratio of
suchsanstalten zu Baden–Baden vom 6–12, September chlorophyll a to chlorophyll b and Douglas-fir seed-
1880. Julius Springer, Berlin.
Booth, J. 1882. Die Naturalisation ausländischer Waldbäume in
Deutschland. Julius Springer, Berlin.
Booth, J. 1890. Die Naturalisation der Douglasfichte ein
Haz- ardspiel? Zeitschrift für Forst und Jagdwesen
22(1):32–51.
Booth, J. 1903. Die Einführung ausländischer Holzarten in die
preussischen Staatsforsten unter Bismarck und anderes.
Julius Springer, Berlin.
Booth, J. 1904. “Grüne” oder “blaue” Douglasfichte? Mitteilun-
gen der Deutschen Dendrologischen Gesellschaft 13:41–42 (in
2nd edition p. 160–161).
Booth, J. 1907a. Die Douglasfichte seit ihrer Einführung nach
Europa (1828–1906). Allgemeine Forst- und
Jagdzeitung. 83:5–10, 45–50, 87–93, 113–118.
Booth, J. 1907b. Die Versuche mit ausländischen Holzarten
in den bayrischen Staatswaldungen.
Forstwissenschaftliche Centralblatt 29:531–541, 598–
606.
Borchers, K. 1951. Folgerungen aus den bisherigen
Anbau- ergebnissen mit fremdländischen Holzarten
im Gebiet des Landes Niedersachsen für die künftige
waldbauli- che Planung. Mitteilungen der Deutschen
Dendrologischen Gesellschaft. 57:69–81.
Borchers, SL, and DA Perry. 1989. Growth and
ectomycor- rhizal formation of Douglas-fir seedlings
grown in soils collected at different distances from
pioneering hard- woods in Southwest Oregon clear
cuts. Canadian Journal of Forest Research 20:712–
721.
Borchert, R. 1976. The concept of juvenility in woody plants.
Acta Horticulturae 56:21–36.
Borchert, R. 1991. Growth periodicity and dormancy, pp.
221–245 in Physiology of Trees, ed. AS Raghavendra.
John Wiley, NY.
Boreau, E. 1938. Recherches anatomiques et expérimentales sur
l’ontogénie des plantules des Pinacées et ses rapports avec
la phylogénie. These de Doctorat, University of Poitiers,
France.
Borer, F. 1982. Zum Wasserhaushalt einer dominierenden
Douglasie in einem Waldbestand. Mitt. Eidg. Anst.
Forstl. Versuchswes. 58(1):5–162.
Bornebusch, CH. 1939. Douglas-fir seed. Quarterly Journal of
Forestry 33:184–186.
Børtnes, G. 1970. Skogreisingen p Vestlandet. (Afforestation
in western Norway.) Tidsskrift for Skogbruk 78:239–
250.
Bouchon, J. 1984. Importance des plantations de Douglas et
épicéa en France. Revue Forestière Française 36(4):254–
258.
Bower, RC, and SD Ross. 1985. Evaluation of ULV sprayer
for seed orchard application of GA 4/7, pp. 103–110, in
New Ways in Forest Genetics, Proceedings Twentieth
Meeting Canadian Tree Improvement Association. Part 2,
ed. F Caron, AG Corrieau, and TJ Boyle.
Bowers, NA. 1942. Cone Bearing Trees of the Pacific Coast.
Whit-
tlesey House, New York.
Boyce, JS. 1932. Decay and other losses in Douglas fir in
western Oregon and Washington. US Department of
Agriculture Technical Bulletin 286. 59 p.
Boyce, JS, and JWB Wagg. 1953. Conk rot of old-growth
Douglas- fir in western Oregon. Bulletin 4, Oregon
Forest Products Laboratory and Oregon State Forestry
Department, Cor- vallis, OR. 96 p.
Boyer, WD. 1966. Long leaf pine pollen dispersal. Forest
Sci- ence 12(3):367–368.
Boyer, WD. 1978. Heat Accumulation: An easy way to
antici- pate the flowering of Southern pines. Journal
of Forestry 76:21–23.
British Columbia Forest Service. 1949. Annual Report of the
ling nutrient status Canadian Journal of Forest Research Lands Service for the year ending December 31, 1948,
23:1655–1662. De- partment of Lands and Forests, Victoria, BC. pp.
Bradbeer, JW. 1988. Seed Dormancy and Germination. Blackie 22–23.
and Son, Ltd., Glasgow, UK. British Columbia Ministry of Forests. 1995a. Bark Beetle
Man- agement Guidebook. Forest Practices Code of BC,
Bradley, AF, WC Fischer, NV Noste. 1992. Fire Victoria, BC. Online:
ecology of the forest habitat types of eastern
http://www.for.gov.bc.ca/tasb/legsregs/fpc/
Idaho and western Wyo- ming. USDA Forest
Service Intermountain Research Sta- tion General
Technical Report INT-290, Ogden, UT. 92 p.
Bramble, WC, and WR Byrnes. 1952. Natural growth
zones in Pennsylvania in relation to varietal testing
of Douglas-fir. Progress Report 81. The
Pennsylvania State College, School of Agriculture.
Bramhall, G. 1966. Permeability of Douglas-fir
heartwood from various areas of growth in British
Columbia. Lumberman 50(1):98, 100, 102.
Bramlett, DL. 1981. Effectiveness of wind pollination in
seed orchards, pp. 10-14 in Pollen Management
Handbook, ed M Franklin. Agriculture Handbook
#587 US Forest Service, Washington, DC.
Brandeis, TJ, M Newton, and EC Cole. 2001.
Underplanted conifer seedling survival and growth
in thinned Douglas- fir stands. Canadian Journal of
Forest Research 31:302–312
Braun, H. 1985. Die Nutzung der Klonselektion als
Züchtungs- methode am Beispiel des
internationalen IUFRO–Doug- lasien–
Herkunftsversuches 1970 auf der Versuchsfläche
Stralsund. Beiträge für die Forstwirtschaft
19(3):133–137.
Braun, H. 1988. Ergebnisse der Hybridzüchtung bei der
Doug- lasie (Pseudotsuga menziesii (Mirb.)
Franco). Beitrage für die Forstwirtschaft 22(1):1–7.
Braun, H, and H Schmiedel. 1985. Ergebnisse der
Anbau- prufung intraspezifischer
Douglasienhybriden unter besonderer
Berucksichtigung der Frostresistenz. Beiträge für
die Forstwirtschaft 19(2):69–73.
Braun, H, and S Weissleder. 1986.
Nachkommenschaftsprü- fung von
Douglasienbeständen der DDR–ein Beitrag zur
Nutzung eigener Reserven. Beiträge für die
Forstwirtschaft 20(4):180–183.
Bravdo, BA, I Levin, and R Assaf. 1992. Control of
root size and root environment of fruit trees for
optimal fruit production. Journal of Plant
Nutrition 15:699–712.
Breidenstein, J, J-C Bastien, and B Roman-Amat. 1990.
Douglas- fir range-wide variation results from the
IUFRO data base. Paper 2.13:1-19, in Proceedings
of the Joint Meeting of Western Forest Genetics
Association and IUFRO Working Parties S2.02-05,
06, 12, and 14. Douglas-fir, Contorta pine, Sitka
spruce, and Abies Breeding and Genetic Resources,
Olympia, WA, August 20-24, 1990. Weyerhaeuser
Co., Federal Way, Centralia, WA.
Brewbaker, JL, and BH Kwack. 1963. The essential role
of cal- cium ion in pollen germination and pollen
tube growth, American Journal of Botany 50 (9,
Oct):859–865.
Bringuel, GJ. 1968. Two new parasites of the Douglas-
fir gall midge, Contarinia oregonensis Foote
(Diptera: Cecido- myiidae). Pan-Pacific
Entomologist 44(4):339–340.
Brink, RA. 1962. Phase change in higher plants and
somatic cell heredity. Quarterly Review of Biology
37(1):1–22.
British Columbia Forest Service. 1936. Forest Research
Review. Forest Research Division, BC Forest
Service, Victoria, BC. pp. 30–31.
British Columbia Forest Service. 1940. Annual Report
of the Lands Service for the year ending December
31, 1939, De- partment of Lands and Forests,
Victoria, BC. pp. 16–17.
292 Douglas-fir: The Genus Pseudotsuga
fpcguide/beetle/betletoc.htm Browning, G. 1985. Reproductive behavior of fruit tree crops and
British Columbia Ministry of Forests. 1995b. Defoliator its implications for manipulation of fruits, pp. 409–425 in
Man- agement Guidebook. Forest Practices Code of BC, Attributes of Trees as Crop Plants. MGR Cannell and JE
Victoria, BC. Online:
http://www.for.gov.bc.ca/tasb/legsregs/fpc/
fpcguide/defoliat/defoltoc.htm
British Columbia Ministry of Forests, Lands, and Natural
Resource Operations. 2014. Okanagan Shuswap Forest
District - Forest Health: Defoliators. Online http://www.
for.gov.bc.ca/dos/programs/forest_health/defoliators.htm.
British Columbia Ministry of Forests. 1984. Forest and
Range Resource Analysis, 1984. Victoria, BC.
Brix, H. 1969. Effect of temperature on dry matter
production of Douglas-fir seedlings during bud
dormancy. Canadian Journal of Botany 47(7):1143–
1146.
Brix, H. 1970. Effects of light intensity on growth of
Western hemlock and Douglas-fir seedlings. Canadian
Forestry Service, Department of Fisheries and Forestry.
Bi-monthly Research Notes 26:34–35.
Brix, H. 1971. Effects of nitrogen fertilization on
photosynthesis and respiration in Douglas-fir. Forest
Science 17:407–414.
Brix, H, and LF Ebell. 1969. Effects of nitrogen
fertilization on growth, leaf area, and photosynthesis
rate in Douglas-fir. Forest Science 15(2):189–196.
Brockley, RP. 1988. The effects of fertilization on the
early growth of planted seedlings: A problem analysis.
FRDA Report 011. Canadian Forest Service, BC
Ministry of For- ests, Victoria, BC. 16 p.
Brodie, JD, and JD Walstad. 1987. Douglas-fir growth and
yield response to vegetation management, pp. 273–
294 in Forest Vegetation Management for Conifer
Production, ed. JD Walstad and PJ Kuch. John Wiley
and Sons, New York. 523 p.
Brodovich, TM. 1964. Investigations of Pseudotsuga
taxifolia
stands in the western Ukraine. Lesnoi Zhurnal 7(4):36–
42. Brodovich, TM. 1967. Oput rasvedenia Pseudotsuga
privivkami
v USSR (Propagation of Pseudotsuga by grafting in
the
Ukrainian Soviet Socialist Republic). Lesnoi Zhurnal,
N.I. 1967. Newsletter of USSR Universities. Translation
No. 55, 1968. The University of British Columbia,
Faculty of Forestry, Vancouver, BC.
Brodovich, TM. 1978. Aklimatizatsiya i adaptatsiya duglasii
tissolistnoi v lesnykh nasazhdeniyakh zapada USSR
(Acclimatization and adaptation of Pseudotsuga
menziesii in forest stands in the western Ukraine).
Lesnoi Zhurnal 21(4):33–36.
Brookes, MH, JJ Colbert, RG Mitchell, and RW Stark,
technical coordinators. 1985. Managing trees and stands
susceptible to western spruce budworm. Canadian
Forestry Service, Headquarters, Ottawa. USDA Forest
Service Technical Bulletin No. 1695. 111 p.
Brookes, MH, RW Stark, and RW Campbell, eds. 1978. The
Douglas-fir Tussock Moth: A Synthesis. USDA Forest
Service, Science and Education Agency, Technical
Bulletin 1585. Washington, DC. 331 p.
Brown, AC, and WA Sinclair. 1981. Colonization and infec-
tion of primary roots of Douglas-fir seedlings by the
ectomycorrhizal fungus Lacarria laccata. Forest
Science 27(1):111–124.
Brown, PC, and S Jones. 1989. Wind risks in the Bay of
Plenty, pp. 20–22 in Workshop on wind damage in New
Zealand exotic forests, ed. AR Somerville, S Wakelin,
and L Whitehouse. New Zealand Forest Research
Institute FRI Bull. No. 146.
Brown, PM, MW Kayes, and D Buckley. 1999. Fire history
in Douglas-fir and coast redwood rests at Point Reyes
Na- tional Seashore, California. Northwest Science 73:205–
216.
Jackson eds. Institute of Terrestrial Ecology, Burr, KE. 1990. The target seedling concept: bud dormancy
National Environmental Research Council Abbots and cold-hardiness, pp. 79–90 in Target Seedling
Ripton, Hunting- don, England. Symposium: Proceeding of the Western Forest Nursery
Browning, JE, and RL Edmonds. 1993. Influence of soil Associations. ed R
alumi- num and pH Armillaria root rot in Douglas-
fir in Western Washington. Northwest Science
67(1):37–43.
Brun Bolte, WR. 1963. Análisis fustal y mecanico di
un pino oregón. Ingeniero Forestal thesis.
Universidad Austral de Chile, Facultad de
Ingeniera Forestal, Instituto de Utilización de
Bosques y Technologia Mecanica de la Madera,
Valdivia, Chile.
Brünig, EF. 1974. Unconventional production programs
for the rehabilitation of gale-damaged private
forests. Forstarchiv 45(10):159–199.
Buch, MV. 1965. Crecimiento juvenil de Pseudotsuga
menziesii y otras especies exóticas en el fundo
“Voipir” de Villarica. Instituto Forestal Santiago,
Boletin Informativo No. 10:60–69.
Buch, MV. 1979. Reactivación de suelos andosolicos
degrados mediante manejo forestal–a causa de la
exploitación del suelo efectuada por los colonos en
el sur de Chile. Mit- teilungen der Bundesanstalt für
Forst- und Holzwirtschaft Hamburg–Reinbek No.
124:2–6.
Buchet, S. 1913. La pretendue heredite des maladies
cryp- togamiques. Bulletin de la Société botanique
de France LIX:754–762.
Buchholz, JT. 1920. Embryo development and
polyembryony in relation to the phylogeny of
conifers. American Journal of Botany 7(4):125–
145.
Bucholz, JT. 1926. Origin of cleavage polyembryony in conifers.
Botanical Gazette 8:55–71.
Buchholz, JT. 1948. Generic and sub-generic distribution of
the Coniferales. Botanical Gazette 110: 80–91.
Buchwald, FN. 1940. Bør nye douglasie–kulturer
anlaegg- es i oeblikket (Should new Douglas-fir
plantations be established at present). Dansk
Skovforenings Tidskrift 25(10):521–527.
Buffam, PE, and NE Johnson. 1966. Tests of Guthion
and Dimethoate for Douglas-fir cone midge
control. Forest Science 12(2) 160–163.
Buffi, R. 1987. Zur Bewirtschaftung der Douglasie
auf der Alpensüdseite. Bündenerwald 40:39–42.
Burdett, AN. 1979. New methods for measuring root
growth capacity: Their value in assessing lodgepole
pine stock quality. Canadian Journal of Forest
Research 9:63–67.
Burdett, AN. 1987. Understanding root growth
capacity: Theoretical considerations in assessing
planting stock quality by means of root growth
tests. Canadian Journal of Forest Research
17:768–775.
Burdett, AN. 1990. Physiological processes in plantation
establishment and development of specifications for
forest planting stock. Canadian Journal of Forest
Research 20:415–427.
Burdett, AN, DG Simpson, and CF Thompson. 1983.
Root development and plantation establishment
success. Plant and Soil 71:103–110.
Burgess, D. 1991. Western hemlock and Douglas-fir
seedling development with exponential rates of
nutrient addition. Forest Science 37:54–67.
Bürgi, A, and C Diez. 1986. Übersicht über den
Exotenanbau in der Schweiz aufgrund einer
Umfrage vom Herbst/ Winter 1984/85.
Schweizer Zeitschrift für Forstwesen
137(10):833–851.
Burr, KE. 1987. Cold Hardiness, Root Growth Capacity,
and Bud Dormancy Testing of Conifer Seedlings. PhD.
Thesis, 105 p. Colorado State University, Fort Collins,
CO.
References 293

Rose, SJ Campbell, and TD Landis. RM-200, USDA Campbell, RK. 1974. Use of phenology for examining prov-
Forest Service, Rocky Mountain Forest and Range enance transfers in reforestation of Douglas-fir. Journal
Experiment Station, Fort Collins, CO. of Applied Ecology 11:1069–1080.
Burr, KE, RW Tinus, SJ Wallner, and RM King. 1989. Rela- Campbell, RK. 1978. Regulation of bud-burst timing by tem-
tionships among cold hardiness, root growth potential perature and photoregime during dormancy, pp. 19–34
and bud dormancy in three conifers. Tree Physiology in Proceedings of the Fifth North American Forest
5(3):291–306. Biology Workshop. ed. CA Hollis, AE Squillace, School
Burr, KE, RW Tinus, SJ Wallner, and RM King. 1990. Com- of Forest Resources, University of Florida, Gainesville.
parison of three cold hardiness tests for conifer seedlings. Campbell, RK. 1979. Genecology of Douglas-fir in a watershed
Tree Physiology 6:351–369. in the Oregon Cascades. Ecology 60(5):1036–1050.
Burzynski, G, and J Gutowski. 1973. Recent results of prov- Campbell, RK. 1986. Mapped genetic variation of Douglas-
enance experiments with Douglas-fir in the Federal fir to guide seed transfer in Southwest Oregon. Silvae
Research Institute in Poland, pp.90–104 in Proceedings Genetica 35(2/3):85–96.
Meeting of IUFRO Working Party on Douglas-fir Campbell, RK. 1987. Biogeographical distribution limits of
Provenances S2.02–05, Sept.3–5, Göttingen, Germany. ed. Douglas-fir in Southwest Oregon. Forest Ecology and
HH Hattemer. Vancouver, BC. Man- agement 18(1):1–34.
Burzynski, G, J Gorczynski, and W Zielinski. 1990. Paper Campbell, RK. 1991. Soils, seed-zone maps, and
2.27 in Proceedings Joint Meeting of the WFGA and physiography: Guidelines for seed transfer of Douglas
IUFRO Working Parties S2.02–05, –06, –12, –14. August fir in southwestern Oregon. Forest Science 37(4):973–
20–24, 1990, Olympia, Washington. Weyerhaeuser Co., 986.
Centralia, WA.
Campbell, RK, and FC Sorensen. 1973. Cold-acclimation
Buszewicz, G, and GD Holmes. 1957. Seven years seed
testing experience with the tetrazolium viability test in seedling Douglas-fir related to phenology and prov-
for conifer species, pp. 142–151 in Report of the enance. Ecology 54(5):1148–1151.
Forestry Research Committee. Great Britain Forestry Campbell, RK, and FC Sorensen. 1978. Effect of test
Commission (London). environ- ment on expression of clines and on
Byrnes, WR, HD Gerhold, and WC Bramble. 1958. Douglas- delimitation of seed zones. Theoretical and Applied
fir Varietal Tests for Christmas Tree Plantations in Genetics 51:233–246.
Pennsylvania. Pennsylvania Agricultural Experiment Campbell, RC, and FC Sorensen. 1979. A new basis for
Station Progress Report 198. charac- terizing germination. Journal of Seed Technology
Cafferata, SL. 1986. Douglas-fir stand establishment over- 4(2):24–32.
view: Western Oregon and Washington, pp. 211–218 in Campbell, RK, and AI Sugano. 1979. Genecology of bud-
Douglas-fir: Stand management for the future, ed. CD burst phenology in Douglas-fir: Response to flashing
Oliver, DP Hanley, and JA Johnson. University of temperature and chilling. Botanical Gazette 140(2):223–
Washington, College of Forest Resources, Institute of 231.
Forest Resources, Contribution 55. Seattle. Campo–Duplan, M. van. 1950. Recherchers sur la
Cahill, JM, TA Snellgrove, and TD Fahey. 1986. The case phylogénie des Abietinées d’après leur grains de
for pruning young-growth stands of Douglas-fir, pp. pollen. Travaux du Laboratoire Forestier Toulouse,
123–131 in Douglas-fir: Stand management for the Tome II, Section 1, vol. 4
future, ed. CD Oliver, DP Hanley, and JA Johnson. :1–182.
University of Washington, College of Forest Cannell, MGR, PM Tabbush, JD Deans, MK
Resources, Institute of Forest Resources, Contribution Hollingsworth, LJ Sheppard, JJ Philipson, and MB
55. Seattle. Murray. 1990. Sitka spruce and Douglas-fir seedlings
Cajander, AK. 1926. Die Entwicklung der Kultur in the nursery and in cold storage: root growth
ausländischer Holzarten in Finnland. Mitteilungen der potential, carbohydrate content, dormancy, frost
Deutschen Den- drologischen Gesellschaft 36:72–75. hardiness and mitotic index. Forestry 63(1):9–27.
Callaham, RZ. 1964. Provenance research: investigation of Cannon, D. 1909. Culture des arbres exotiques en Sologne. Nou-
genetic diversity associated with geography. Unasylva velle édition (Original not seen, cited from Hickel 1922).
18(2–3):40–50.
Carlson, C. 1980. Kraft mill gases damage Douglas-fir in
Callaham, RZ. 1970. Introduction: Forest Tree Breeding. Una- western Montana. European Journal of Forest
sylva 24(2-3):97–98. Pathology 10(2/3):145–151.
Callaham, RZ. 2013. Pinus ponderosa: geographic races Carlson, CE, and JE Dewey. 1971. Environmental pollution
and subspecies based on morphological variation. by fluorides in Flathead National Forest and Glacier
Research Paper PSW-RP-265. USDA Forest Service, National Park. USDA Forest Service, Division of State
Pacific Sou- thwest Research Station, Albany, CA. 53 and Private Forestry, Forest Insects and Diseases
p.
Branch, Missoula, MT. 57 p.
Camefort, H. 1962. L’organisation du cytoplasme dans l’oos-
phere et la cellule centrale du Pinus laricio poir (var. Carlson, CE, and CJ Gilligan. 1983. Histological
aus- triaca). Annales des Sciences Naturelles Botanique et differentiation among abiotic causes of conifer needle
Biologie Vegetale 12:269–291. necrosis. Research Paper INT-298, USDA Forest Service
Intermountain Range and Experiment Station, Ogden,
Camefort, H. 1967. Observations sur les mitochondries et UT. 17 p.
les plastes de la cellule centrali et de l’oosphere di Laris
decidua Mill (Laris europa DC). Comptes Rendus de Carlson, CE, CC Gordon, and CC Gilligan. 1979. The
l'Aca- démie des Sciences Paris 265:1293–1296. relation- ship of fluoride to visible growth/health
characteristics of Pinus monticule, Pinus contorta, and
Camm, EL, DC Goetze, SN Silim, and DP Lavender. 1994. Pseudotsuga menziesii. Fluoride 12(1):9–17.
Cold storage of conifer seedlings: An update from the
British Columbia perspective. The Forestry Chronicle Carlson, WC. 1978. The use of periodic moisture stress to
70(3):311–316. induce vegetative bud set in Douglas-fir seedlings. In-
ternational Plant Propagators’ Society 28:49–58.
Campbell, RK. 1964. Recommended traits to be improved in
a breeding program for Douglas-fir. Weyerhaeuser Carlson, WC, and CL Preisig. 1981. Effects of controlled
Company Forest Research Note No 57. 19 p. release fertilization on shoot and root development of
Campbell, RK. 1972. Genetic variability in juvenile height- Douglas-fir seedlings. Canadian Journal of Forest
growth of Douglas-fir. Silvae Genetica 21:126–129. Research 11:230–242.
Carlson, WC, WD Binder, CO Feenan, and CL Preisig. 1980.
Changes in mitotic index during onset of dormancy
in Douglas-fir seedlings. Canadian Journal of Forest
Research 10:371–378.
294 Douglas-fir: The Genus Pseudotsuga
Carmichael, RL. 1957. Relation of seeding date to Chappell, HN, SAY Omule, and SP Gessel. 1992. Fertilization
germination in Coastal Northwest Forests: Using response information
of Douglas-fir seed. Northwest Science 31:177–182.
Carrasco, MM. 1954. Primeros resultados de reforestación
con coniferas exóticas en las Zonas Pre–Cordilleranas
de la Provincia de Valdivia, pp. 12–14 in Bosques y
Maderas. Primer Congreso Forestal Maderero
Nacional. Corpo- ración Chilena de la Madera,
Santiago, de Chile.
Carter, R. 1992. Diagnosis and interpretation of forest stand
nutrient status, pp. 90–97 in Forest fertilization:
sustaining and improving nutrition and growth of western
forests, ed. HN Chappell, GF Weetman, and RE Miller.
Institute of Forest Resources Contribution 73, College of
Forest Resources, University of Washington, Seattle.
302 p.
Carvalho, A de. 1965. Contribuiçãp para o estudo das
principais madeiras de resinosa introduzidas no perimetro
florestal de Manteigas (Serra da Estrella). Estudos e
Divulgação Tecnica, Secretaria de Estado da Agricultura,
Direcção–Geral dos Servicos Florestais e Aquícolas,
Lisbon. (Original not seen, cited from Forestry Abstracts
31:6211, 1970).
Carvalho, A de. 1990. Debilidade adaptativa de exóticas o
exemplo da Pseudotsuga, in Reprint of paper given at
the Second Congresso Florestal Nacional in Porto,
Portugal, Nov. 7–10, 1990.
Castellano, MA. 1994. Current status of out planting studies
using ectomycorrhizal-inoculated forest trees, pp. 261–
281 in A Reappraisal of Mycorrhizae in Plant Health.
American Phytopathological Society, ed. FL Pleger and
B Linderman, St. Paul, MN.
Castellano, MA, and JM Trappe. 1985. Ectomycorrhizal
forma- tion and plantation performance of Douglas-fir
nursery stock inoculated with Rhizopogon spores.
Canadian Journal of Forest Research 15:613–617.
Chalupka W. 1997. Carry-over effect of gibberellins (GA4/7)
and ringing on female flowering in Norway spruce (Pi-
cea abies (L) Karst) seedlings. Annals of Forestry
Science 54:237–341.
Chamberlain, CJ. 1935. Gymnosperms: Structure and
Evolution.
University of Chicago Press, Chicago, IL.
Chandler, WH. 1957. Deciduous Orchards. 3rd ed., Lea
and Febiger, Philadelphia, PA.
Chanway, CP. 1997. Inoculation of tree roots with plant
growth promoting soil bacteria: An emerging technology
for reforestation. Forest Science 43(1):99–112.
Chanway CP, and FB Holl. 1991. Biomass increase and as-
sociative nitrogen fixation of mycorrhizal Pinus contorta
seedlings inoculated with a plant growht promoting
Bacillus strain. Canadian Journal of Botany 69(3):507–
511.
Chanway CP, and FB Holl. 1992. Influence of soil biota on
Douglas-fir (Pseudotsuga menziesii) seedling growth:
the role of rhizosphere bacteria. Canadian Journal of
Botany 70(5):1025–1031.
Chanway CP, and FB Holl. 1994. Ecological growth
response specificity of two Douglas-fir ecotypes
inoculated with coexistent beneficial bacteria. Canadian
Journal of Botany 72(5):582–586.
Chanway, CP, Radley, RA, and FB Holl. 1991a. Inoculation
of conifer seed with plant growth promoting Bacillus
strains causes increased seedling emergence and
biomass. Soil Biology and Biochemistry 23:575–590.
Chanway, CP, R Tukkington, and FB Holl. 1991b.
Ecological implications specifity between plants and
rhizosphere mi- croorganisms. Advances in Ecological
Research 21:121–169.
Chappell, HN, GF Weetman, and RE Miller, eds. 1992. For-
est Fertilization: Sustaining and Improving Nutrition
and Growth of Western Forests. Institute of Forest
Resources Contribution #73, College of Forest
Resources University of Washington, Seattle. 302 p.
in developing stand-level tactics, pp. 1098–1113 in Flower induc-
Forest Fertilization: Sustaining and Improving tion in Douglas -fir (Pseudotsuga menziesii [Mirb.] Franco)
Nutrition and Growth of Western Forests, ed. HN by fertilizer and water stress. Part 1: Changes in free
Chappell, GF Weetman, and RE Miller, Institute of amino acid composition in needles, pp. 43–54 in Sexual
Forest Resources Contribution 73, College of Forest Reproduction of Conifers: Symposium Proceedings,
Resources University of Washington, Seattle. 302 p. Academy of Science USSR, Siberian Division, Nauka,
Charbon, D. 1991. Les insolites arbres exotiques dans les forêts 43–54.
Lausannoises. Les cahiers de la forêt Lausannoise 7.
Charpentier, JP, and M Bonnet-Masimbert. 1983.
Influence d'une réhydratation préalable sur la
germination in vitro du pollen de douglas
(Pseudotsuga menziesii) après conser- vation.
Annals of Forest Science 40(3):309–317.
Chatthai, M, and S Misra. 1998. Sequence and
expression of embryogenesis-specific cDNAs
encoding 2S seed storage proteins in Pseudotsuga
menziesii [Mirb.] Franco. Planta 206(1):138–145.
Chen, HYH. 1997. Interspecific responses of planted
seedlings to light availability in interior British
Columbia: survival, growth, allometric patterns, and
specific leaf area. Cana- dian Journal of Forest
Research 27, 1383–1393.
Chen, HYH, and K Klinka. 1997. Light availability and
pho- tosynthesis of Pseudotsuga menziesii seedlings
grown in the open and in the forest understory. Tree
Physiology 17(1):23–29.
Chen, HYH, K Klinka, and GJ Kayahara. 1996. Effects
of light on growth, crown architecture, and specific
leaf area for naturally established Pinus contorta
var. latifolia and Pseudotsuga menziesii var. glauca
saplings. Canadian Journal of Forest Research
26(7):1149–1157.
Chen, Z, TE Kolb, and K Clancy. 2002. The role of
monoterpenes in resistance of Douglas-fir to
western spruce budworm defoliation. Journal of
Chemical Ecology 28(5):897–920.
Chen, ZY, RK Scagel, and J Maze. 1986. A study of
morphologi- cal variation in Pseudotsuga menziesii in
Southwestern Brit- ish Columbia. Canadian Journal
of Botany 64(8):1654–1663.
Cheng, WC, and LK Fu. 1978. Gymnospermae. Vol. 7
in Flora of China. Flora Rei publicae Popularis
Sinicae: delectis florae Rei publicae Sinicae
agendae Academiae Sinicae edita, Tomus 7.
Academy Publishing House, Peking, China.
Chengde, C. 1981. A brief introduction to the Chinese species
of the genus Pseudotsuga. Davidsonia 12(1):15–17.
Chesnoy, L. 1973. Sur l’origine paternelles des organites
de proembryon du Chamaecyparis lawsonia A.
Murr (Cupres- sacées). Carylogia 25 Suppl: 223–
232.
Childs, TW. 1960. Drought effects on conifers in the
Pacific North- west, 1958–1959. Research Note
PNW-182, USDA Forest Service, Pacific
Northwest Forest and Range Experiment Station,
Portland, OR. 5 p.
Childs, TW. 1961. Delayed killing of young-growth
Douglas-fir by sudden cold. Jourbal of Forestry
59(6):452–453.
Ching, KK. 1959. Hybridization between Douglas-fir
and bigcone Douglas fir. Forest Science 5(3):246–
254.
Ching, KK. 1965. Early Growth of Douglas-fir in a
Reciprocal Planting. Research Paper 3. Oregon
Forest Research Labo- ratory, Oregon State
University, Corvallis.
Ching, KK, and DN Bever. 1960. Provenance study of
Douglas- fir in the Pacific Northwest region. I.
Nursery perfor- mance. Silvae Genetica 9(1):11–17.
Ching, KK, and PN Hinz. 1978. Provenance study of
Douglas- fir in the Pacific Northwest region. III.
Field performance at age twenty years. Silvae
Genetica 27(6):229–233.
Ching, KK, and DP Lavender. 1960. Unpublished
data. Ching, KK, TM Ching, and DP Lavender. 1973.
References 295
Scottish Forestry 42(l):21–32.
Ching, TM. 1959. Activation of germination in Douglas-fir Chylarecki, H. 1976. Badania nad daglezją w Polsce w różnych
seed by hydrogen peroxide. Plant Physiology 34(5):557–
563.
Ching, TM. 1961. Fatty acids in Douglas-fir seeds. Plant
Physi- ology 36(supplement):XLV.
Ching, TM. 1963a. Change of chemical reserves in
germinating Douglas-fir seed. Forest Science 9(2):226–
231.
Ching, TM. 1963b. Fat utilization in germinating Douglas-fir
seed. Plant Physiology 38(6):722–728.
Ching, TM. 1965. Metabolic and ultrastructural changes in
germinating Douglas-fir seeds. Plant Physiology 40
(sup- plement): viii–ix.
Ching, TM. 1966. Compositional changes of Douglas-fir
seeds
during germination. Plant Physiology 41(8):1313–1319.
Ching, TM. 1968. Intracellular distribution of lipolytic activ-
ity on female gametophytes of germinating Douglas-
fir
seeds. Lipids 3(6):482–488.
Ching, TM. 1973a. Biochemical aspects of seed vigor. Seed
Science and Technology 1:73–88.
Ching, TM. 1973b. Adenosine triphosphate content and
seed
vigor. Plant Physiology 51:400–402.
Ching, TM, and KK Ching. 1959. Extracting Douglas-fir pol-
len and effects of GA on its germination. Forest Science
5(1):74–80.
Ching, TM, and KK Ching. 1962. Physical and physiological
changes in maturing Douglas-fir cones and seeds. Forest
Science 8(1):21–31.
Ching, TM, and KK Ching. 1973. Energy status in dormant
and non-dormant seeds, pp. 13–19 in lnternational
Sympo- sium on Dormancy in Trees. International Union
of Forest Research Organizations, Polish Academy of
Sciences, Rorah, Poland.
Ching, TM, and KK Ching. 1976. Rapid viability tests and
aging study of some coniferous pollen. Canadian
Journal of Forest Research 6:516–522.
Ching, TM, and SC Fang. 1963. Utilization of labeled
glucose in developing Douglas-fir seed cones. Plant
Physiology 38(5):551–554.
Ching, TM, and LA Jensen. 1958. Sampling experiments of
coniferous tree seeds for testing. Proceedings of the As-
sociation of Official Seed Analysts 49(1):49–55.
Ching, TM, and LA Jensen. 1959. Sampling Experiments of
Coniferous Tree Seeds for Testing. Technical Paper
1237. Oregon Agricultural Experiment Station,
Corvallis, OR.
Ching, TM, and MC Parker. 1958. Hydrogen peroxide for
rapid viability tests of some coniferous tree seeds.
Forest Science 4(2):128–134.
Ching, TM, and I Schoolcraft. 1968. Physiological and
chemical differences in aged seeds. Crop Science 8:407.
Ching TM, and WH Slabaugh WH. 1966. X-ray diffraction
analysis of ice crystals in coniferous pollen. Cryobiology
2(6):321–327.
Ching TM, M Ranzoni, and KK Ching. 1975. ATP content
and pollen germinability of conifers. Plant Science
Letters 4:331–333.
Christiansen, H. 1969a. On the pollen grain and the fertiliza-
tion mechanism of Pseudotsuga menziesii. Silvae
Genetica 18:97–104.
Christiansen, H. 1969b. On the germination of pollen of
Larix and Pseudotsuga on artificial substrate, and on
viability tests of pollen of coniferous forest trees.
Silvae Genetica 18:104–107.
Christiansen, H. 1972. On the development of pollen and
the fertilization mechanisms of Larix and Pseudotsuga
menziesii. Silvae Genetica 21:166–174.
Christie, JM. 1988. Levels of production class of Douglas-fir.
in Douglas-fir sheets. Tree Physiology 26(10):1369–
warunkach ekologicznych (Research of Douglas 1375.
fir in Poland in various ecological conditions). Coaz, J. 1897. Anbau der Douglasia. Schweizerische Zeitschrift
Arboretum Kór- nickie 21: 15-124 (in Polish). für Forstwesen 48:98–100.
Chylarecki, H. 2005. Douglas fir in Polish forests. Coaz, J. 1905. Dendrologische Leistungen in der Schweiz.
Production potential, ecological requirements Mitteilungen Deutsche Dendrologische Gessellschaft 14:51–
biology. Mitteilungen aus der Forschungsanstalt 52. Cochran, PH. 1979. Gross yields for even-aged stands of
fur Waldokologie und Forst- wirtschaft
Rheinland-Pfalz (55/05):37–134. Douglas-
fir and white or grand fir east of the Cascades in Oregon and
Ciancio, O, A Eccher, and G Gemignani. 1980.
Eucalitti, pino insigne, douglasia e altre specie
forestali a rapido accres- cimento. Legno e
Cellulosa. Italia Agricola 117(1):190–218.
Cieslar, A. 1898. Vergleichende Studien über Zuwachs
und Holzqualitaet von Fichte und Douglastanne.
Centralblatt für das gesamte Forstwesen
24(8/9):355–372.
Cieslar, A. 1901. Über Anbauversuche mit
fremdländischen Holzarten in Österreich.
Centralblatt für das gesamte Forst- wesen 27:101–
116, 150–175, 196–209.
Cieslar, A. 1920. Die grüne Douglastanne (Pseudotsuga
Doug- lasii (Carr.) im Wienerwalde. Centralblatt
für das gesamte Forstwesen 46(1/2):3–26.
Cissel, JH, FJ Swanson, GE Grant, DH Olson, SV
Gregory, SL Garman, LR Ashkenas, MG Hunter,
JA Kertis, JH Mayo, MD McSwain, SG Swetland,
KA Swindle, and DO Wallin. 1998. A landscape
plan based on historical fire regimes for a managed
forest ecosystem: the Augusta Creek study. USDA
Forest Service, General Technical Report PNW-
GTR-422, Pacific Northwest Research Station,
Portland, OR. 82 p.
Clancy, KM. 1992a. Response of western spruce
budworm (Lepidoptera: Tortricidae) to increased
nitrogen in arti- ficial diets. Environmental
Entomology 21:331–344.
Clancy, KM. 1992b. The role of sugars in western
spruce budworm nutritional ecology. Ecololgical
Entomology 17:189–197.
Clancy, KM, Z Chen, and TE Kolb. 2004. Foliar
nutrients and induced susceptibility: genetic
mechanisms of Douglas- fir resistance to western
spruce budworm defoliation. Canadian Journal of
Forest Research 34:939–949.
Clark, EC, JH Schenk, and DL Williamson. 1963. The
cone- infesting moth Barbara colfaxiana as a pest
of Douglas-fir in Northern Idaho. Annals of the
Entomological Society of America 56:246–250.
Clark, MB. 1958. Studies on seed dispersal of Douglas-fir.
BC Forest Service Forest Research Review 57–58:19–
20.
Clark, MB. 1969. Direct Seeding Experiments in the
Southern Interior Region of British Columbia.
Research Note 49. Brit- ish Columbia Ministry of
Forests, Department of Lands, Forests, and Water
Research, Victoria, BC. 9 p.
Clear, T. 1951. Douglas-fir in Co. Wicklow. Irish Forestry
8(1):8–18.
Cleary, BD. 1969. Water stress measurements and their
ap- plication to forest regeneration, pp. 29–32 in
Proceedings of the Western Forestry and
Conservation Association, Western Reforestation
Coordinating Committee, Portland OR.
Cleary, BD, RD Greaves, and RK Hermann. 1978.
Regenerating Oregon’s forests. A guide for the
regeneration forester. Oregon State University
Extension Service, Corvallis. 286 p.
Climent JM, MA Prada, LA Gil, and JA Pardos. 1997.
Increase of flowering in Pinus nigra Arn subsp
salzmannii (Dunal) Franco by means of
heteroplastic grafts. Annals of Forest Science
54(2):145–153.
Cline, M, M Yoders, D Desai, C Harrington, and W
Carlson. 2006. Hormonal control of second flushing
296 Douglas-fir: The Genus Pseudotsuga
Washington. USDA Forest Service, Research Paper Corbineau, F, J Bianco, G Garello, and D Come. 2002. Breakage
PNW- of Pseudotsuga menziesii seed dormancy by cold treatment as
263. Pacific Northwest Forest and Range Experiment related to changes in seed AbA sensitivity and AbA levels.
Station, Portland, OR. 17 p. Physiologia Plantarum 114:313–319.
Cogshall, AS. 1928. Food habits of deermice of the genus Cordell, CE and Marx, DH 1994. Effects of nursery cultural
Peromyscus in captivity. Journal of Mammology 9:217–
218.
Cokl, M. 1965. Rast zelene duglazije v Slovenje. Zbornik
Insti- tuta za gozdno in lesno gospodarstvo. Slovenije st.
4:139–187. Ljubljana.
Cole, EC, WC McComb, M Newton, JP Leeming, and CL
Chambers. 1998. Response of small mammals to clear-
cutting, burning and glyphosate application in the
Oregon coast range. Journal of Wildlife 62(4):1207–
1216.
Coleman, MD, C Bledsoe, and BA Smit. 1990. Root
hydraulic conductivity and xylem sap levels of zeatin
riboside and abscisic acid in ectomycorrhizal Douglas
fir seedlings. New Phytologist 115:275–284.
Comerford, NB, WG Harris, and D Lucas. 1990. Release of
nonexchangeable potassium from a highly-weathered,
forested Quartzipsamment. Soil Science Society of
America Journal 54:1421–1426.
Condrashoff, SF. 1968. Biology of Steremnius carinatus
(Coleop- tera: Curculionidae), a reforestation pest in
coastal British Columbia. The Canadian Entomologist
100(4):386–394.
Condrashoff, SF. 1969. Steremnius carinatus (Boheman), a
wee- vil damaging coniferous seedlings in British
Columbia. Government of Canada, Information Report
BC-X-017, Department of Forestry and Rural
Development, Forest Research Laboratory, Victoria, BC.
6 p.
Contreras, CC, and NR Peters. 1982. Indices de sitio para
pino oregón en la Provincia de Valdivia y sus relaciones
con sitios para pino insigne, pp. 98–109 in Proceedings
of the workshop on evaluating the productivity of forest
sites, 22–24 April, 1982. JE Schlatter, compiler.
Universidad Austral de Chile, Facultad de Ciencias
Forestales, Valdivia. (Original not seen, cited from
Forestry Abstracts 47:2727, 1986.)
Contreras, JM, and GB Smith. 1973. Estudio preliminar de
incre- mento y rendimiento de pino oregón (Pseudotsuga
menziesii) en la región sur de Chile. Ingeniero Forestal
thesis. Univer- sidad Austral de Chile, Facultad de
Ingenieria Forestal, Instituto de Manejo y Economia
Forestal, Valdivia, Chile.
Cooley, RA. 1908. The Douglas spruce cone moth. Montana
Agricultural Experiment Station Bulletin 70: 125–130.
Cooley, SJ. 1982. Cone and Seed Disease. Tree Seed Paper
8. USDA Forest Service Pacific Northwest Region, Port-
land, OR.
Cooley, SJ. 1985. Top Sleight of Douglas-fir Seedlings
Fungicide Trials in Five Pacific Northwest Forest
Nurseries, First Years Results. USDA Forest Service,
Pacific Northwest Region Forest Pest Management. 14 p.
Copeland, LO. 1976. Principles of Seed Science and
Technology.
Burgess Publishing Co. Minneapolis, MN.
Copeland, LO, and MB McDonald. 2001. Seed Science and
Technology. 4th Edition. Kluwer Academic Publishers.
Dor- drecht, The Netherlands.
Copes, DL. 1978. A genetic analysis of amino-peptidase and
peroxidaseisoenzymes in Douglas-fir trees and seedling
progeny. Canadian Journal of Forest Research 9:189–192.
Copes, DL. 1983. Failure of grafted Douglas-fir planted at
Monterey. California Tree-Planters’ Notes 34(3):9–10.
Copes, DL, and NC Vance. 2000. Effects of Water
Suspension and Wet-Dry Cycling on Fertility of Douglas-
Fir Pollen. Research Note PNW-RN-527, Pacific
Northwest Research Station, USDA Forest Service
Portland, OR.
practices on management of specific root disease. Forest Pathol- ogy 36:372–384.
ectomycorrhizal on bare root tree seedlings, pp. Cummings, JC. 1962. The Effect of the Microflora of Douglas-fir
133–151 in Mycorrhizae and Plant Health, ed. FL Seed on the Damping Off organism, Fusanium oxysporum
Pfleger and RG Linderman, The American schl.
Phytopthological Society, St. Paul, MN. 344 pp.
Cordell, CE, JH Owen, and DH Marx. 1987.
Mycorrhizae nurs- ery management for improved
seedling quality and field performance, pp. 105–115
in Proceedings, Intermountain Forest Nursery
Association, Oklahoma City, OK.
Cornelius, J. 1994. Heritabilities and additive genetic
coef- fients of variation in forest trees. Canadian
Journal of Forest Research 24:372–379.
Costello, DF. 1954. Vegetation zones in Colorado, pp.
iii–x in Manual of the Plants of Colorado, ed. HD
Harrington. Sage Books, Denver, CO.
Coulson, RN, and RT Franklin. 1970. The biology of
Dioryctria amatiela (Lepidoptera Phycitidae. The
Canadian Entomolo- gist 102:679–683.
Coutinho, AXP. 1936. Esbôco de uma flora Lenhosa Portu-
guesa. Direcção–Geral dos Servicos Florestais e
Aquícolas, vol. 3, Tomo I:334–340.
Coutts, MP. 1987. Developmental processes in tree root
sys- tems. Canadian Journal of Forest Research
17:761–767.
Coutts, MP, and JJ Philipson. 1976. The influence of
mineral nutrition on the root development of trees.
Journal of Experimental Botany 27:1102–1111.
Crabbé, J, and P Barnola. 1996. A new conceptual
approach to bud dormancy in woody plants, pp. 83–
113 in Plant Dormancy, Physiology, Biochemistry
and Molecular Biology, ed. GA Lang. CAB Intl.,
Wallingford, UK.
Craib, IJ, 1939. Thinning, pruning and management
studies on the main exotic conifers grown in South
Africa. Science Bulletin 196. Department of
Agriculture and Forestry, Union of South Africa.
Craib, WG. 1919. A new Chinese Pseudotsuga. Royal
Botanical Garden Edinburgh Notes 60(11):189–190;
pl. 160–161.
Cristofolini, F. 1968. Primi rilievi sulla Douglasia in Liguria.
I’Italia Forestale e Montana 23(3):117–131.
Critchfield, WB. 1984. Impact of the Pleistocene on the
ge- netic structure of North American conifers, pp.
70–131 in Proceedings 8th North American Forest
Biology Workshop, ed. RM Lanner, Department of
Forest Resources, Utah State University, Logan,
UT.
Critchfield, WB, and GL Allenbaugh. 1969. The
distribu- tion of Pinaceae in and near northern
Nevada. Madroño 20(1):12–26.
Crook, RW, and WE Friedman. 1992. Effects of pollen
tube number and archegonium number on
reproduction in Douglas-fir: significance for seed
orchard management. Canadian Journal of Forest
Research 22:1483–1488.
Crosby, CR. 1909. On certain seed-infesting chalcis-flies.
College of Agriculture, Department of Entomology,
Bulletin 265, Cornell University, pp. 367–388.
Crosby, CR. 1913. A Revision of the North American
Species of Megastigmus Dalman. Annals of the
Entomological Society of America 6:155–170.
Crouch, GL, and MA Radwan. 1972. Arasin in Endrin
Treatments to Protect Douglas-fir Seed From Deer
Mice. USDA Forest Service Research Paper PNW-
RP-136. Pacific Northwest Forest and Research
Experiment Station, Portland, OR.
Crozier, JD. 1908. The Douglas-fir as a commercial timber
tree.
Royal Scottish Arboricultural Society Transactions 21(1):31–
40
+ pl. VII.
Cruickshank, MG, D Lejour, and DJ Morrison. 2006.
Traumatic resin canals as markers of infection
events in Douglas-fir roots infected with Armillaria
References 297
Das, ES, and B Chand. 1958. Experimental afforestation of bare,
BS thesis, University of British Columbia, Vancouver, dry, and rocky slopes in high hills. Indian Forester
BC. Curchod, Ed. Essais d’acclimatation d’essences exotiques
faits
dans les forets de la commune de Lausanne. Jouru.
Forest.
Suisse, Berne, 52, 1901, (142, 162, 175).
Curtis, JD. 1964. Roots of a Ponderosa Pine. USDA Forest
Service Research Paper INT. 9, Intermountain Forest and
Range Experiment Station. Ogden, UT.
Curtis, RO, and DD Marshall. 1986. Levels of growing stock
cooperative study in Douglas-fir. Report #8– The LOGS
study: Twenty-year results. USDA Forest Service Research
Report PNW 356, Pacific Northwest Forest and Range
Experi- ment Station, Portland, OR. 173 p.
Curtis, RD, and DL Reukema. 1970. Crown development and
site estimates in a Douglas-fir plantation spacing test.
Forest Science 16(3):287–301.
Curtis, RO, FR Herman, and DJ DeMars. 1974. Height
growth and site index for Douglas-fir in high-
elevation forests of the Oregon-Washington Cascades.
Forest Science 20(4):307–316.
Curtis, RO, GW Clendenen, DL Reukema, and DJ DeMars.
1982. Yield tables for managed stands of coast Douglas-
fir. USDA Forest Service, General Technical Report
PNW-135. Pacific Northwest Forest and Range
Experiment Station, Portland, OR. 182 p.
Curtis, RD, DS DeBell, CA Harrington, DP Lavender, JB St
Clair, JC Tappeiner, and JD Walstad. 1998. Silviculture
for Multiple Objectives in the Douglas-fir Region.
USDA Forest Service General Technical Report, PNW-
GTR-435, Pacific Northwest Forest and Range
Experiment Station, Portland, OR.
Dallimore, W. 1932. W Lobb and J Jeffrey. Forestry 6(1):5–8.
Dallimore, W, and BA Jackson. 1948. A Handbook of
Coniferae.
Ed. 3. Edward Arnold and Co., London.
Danckelmann, B. 1884. Die Ergebnisse der in den Preus-
sischen Staatsforsten ausgeführten Anbauversuche mit
fremdländischen Holzarten. Zeitschrift für Forst-und
Jag- dwesen 16(6):289–315.
Daniel, TW. 1981. The middle and southern Rocky Mountain
region, pp. 277–340 in Regional silviculture of the
United States, ed. John W. Barrett. John Wiley, New
York.
Danielson, HR, and Y Tanaka. 1978. Drying and storing
strati- fied ponderosa pine and Douglas-fir seeds. Forest
Science 24(1):4–16.
Danks, HV. 1987. Insect Dormancy: An Ecological
Perspective. Biological Survey of Canada (Terrestrial
Arthropods), Ottawa. ix + 439 p.
Daoudi, EH, and M Bonnet-Masimbert. 1998. Polyamines
conjuguées et différenciation florale chez le sapin de
Douglas (Pseudotsuga menziesii (Mirb.) Franco).
Canadian Journal of Botany 76(5):782–790.
Daoudi, EH, M Bonnet-Masimbert, and J Martin-Tanguy.
1991. Évolution des polyamines dans les bourgeons et
les rameaux de Pseudotsuga menziesii (Mirb) Franco
à la suite du passage de l’état végétatif à l’état floral.
Annals of Forest Science 48(4):367–376.
Daoudi, EH, P Doumas, and M. Bonnet-Masimbert. 1994.
Changes in amino acids and polyamines in shoots and
buds of Douglas-fir trees induced to flower by
nitrogen and gibberellin treatments. Canadian Journal
of Forest Research 24(9):1854–1863.
Daoudi, EH, Plour, A de, and Beaulieu, O. 1995.
Indications of crown size and maturing of flower
production and sex expression in Picea glauca treated
with gibberellin A and
H. Tree Physiology 15:441–445.
Darwin, C. 1859. On the Origin of Species. The Folio
Society,
London, 2006
provenances, S2.02–06 Lodge- pole pine provenances,
84(7):402–406. S2.02–12 Sitka spruce provenances, S2.02–14 Abies
provenances, Vancouver, British Columbia, 1978. British
Daubenmire, R. 1960. A seven-year study of cone Columbia Ministry of Forests, Information Services
produc- tion as related to xylem layers and Branch, Victoria, BC.
temperature in Pinus ponderosa. The American
Midland Naturalist 64:187–193.
Davidson, A, and GL Moxley. 1923. Flora of Southern California.
Times-Mirror Press, Los Angeles.
Davies, PJ, ed. 1987. Plant Hormones and Their Role in
Plant Growth and Development. Martinus Nijhoff
Publ., Dor- drecht, The Netherlands.
Dawkins, 2009. The Greatest Show on Earth. Simon and Schus-
ter, Inc. New York.
Day, RJ. 1982. Evaluating root regeneration potential of
bare- root nursery stock, pp. 83–96 in Proceedings
of the 1981 Intermountain Nurserymen’s
Association Meeting, August 11–13, 1981,
Edmonton, Alberta, comp. RF Huber. Environ- ment
Canada, Forest Service Information Report NOR-
X-241, Northern Forestry Research Centre,
Edmonton, AB.
Day, WR. 1955. The place of a species in the forest,
with special reference to Western North-
American species of conifer used in Britain.
Forestry 28:33–47.
De Champs, J. 1997a. La conduite des peuplements de
Douglas, pp. 1091–209 in Le Douglas, ed. J. De
Champs. Association Forêt-Cellulose (AFOCEL),
Paris.
De Champs, J. 1997b. Le Douglas en Europe et en
France, pp. 35–51 in Le Douglas, ed. J. De Champs.
Association Forêt-Cellulose (AFOCEL), Paris.
De Champs, J. 1997c. La Plantation, pp. 128–129 In: Le Douglas,
J De Champs, coord. AFOCEL, Paris, France.
De Champs, J, JL Ferron, D Michaud, and N Savatier.
1982. Lecons a tirer dela tempête des 6–8
Novembre, 1982. AFOCEL Annales de recherches
sylvicoles 1982:4–101.
De Hoogh, J. 1924. De groene Douglas in Nederland
(with Engl. summary: The Douglas fir in the
Netherlands). Mededelin- gen van het
Rijksbosbouwproefstation, Deel II, Aflevering
1:13–114.
De Hoogh, J. 1925. Bijdrage tot de kennis van de
groei van Pseudotsuga taxifolia Britt in
Nederland, in verband mit zijn beteekerris voor
der Nederlandsche Boschbouw.- Wageningen.
De Groot, P, JJ Turgeon, and GE Miller. 1994. Status of
cone and seed insect pest management in Canadian
seed or- chards. Forestry Chronicle 70(8):745–761.
De Keijzek, S, and RK Hermann. 1965. Effect of
charcoal on germification of Douglas-fir seed.
Northwest Science 40(4):155–163.
De Matos Malavasi, M, TM Ching, and DP Lavender.
1986. Stratifying, partially redrying, and storing
Douglas-fir seeds: biochemical responses. Annals
of Forest Science 43:35–48.
De Matos Malavasi, M, SG Stafford, and DP
Lavender. 1985. Stratifying, partially redrying,
and storing Douglas-fir seeds: effects on growth
and physiology during germina- tion. Annals of
Forest Science 42:371–384.
de Vecchi, E. 1969. Evolution and importance of land races in
breeding. FAO–FO–FTB–69–1045:1263–1278.
de Vecchi, E. 1973. IUFRO Douglas-fir plantations
made by the Istituto Nationale per Piante da Legno
“G. Piccarolo” of Turin, in spring 1970, pp. 188–194
in Proceedings Meeting of IUFRO Working Party on
Douglas-fir provenances S2.02–05, September 3–5,
Göttingen, Germany, ed. HH Hattemer.
de Vecchi, E. 1978. IUFRO Douglas-fir plantation in Turin,
Italy, pp. 347–348 in Vol. 1: Background papers and
Douglas-fir provenances. Proceedings IUFRO joint
meeting of Working Parties S2.02–05 Douglas-fir
298 Douglas-fir: The Genus Pseudotsuga
de Vries, PG. 1961. Een onderzoek naar de productiviteit man- aging the impact of soil compaction on forest lands.
van verschillende douglas–herkomsten in Nederland. OSU Extension Service and OSU Department of Forest
Mededelingen van de landbouwhogeschool to Wageningen,
Nederland 6(13):1–40.
de Vries, SMG. 1990. Use and improvement of Douglas-fir
in the Netherlands. Paper 4.37 in Proceedings Joint
Meeting of Western Forest Genetics Association (WFGA)
and IUFRO Working Parties S2.02–05, 06, 12 and 14,
Olympia, Wash-
ington, August 20–24, 1990.
Deans, JD, C Lundberg, PM Tabbush, MGR. Cannell, LJ
Sheppard, and MB Murray. 1990. The influence of
desic- cation, rough handling and cold storage on the
quality and establishment of Sitka spruce planting stock.
Forestry 63(2):129–141.
DeBell, DG, CA Harrington, and J Shumway. 2002.
Thinning shock and response to fertilizer less than
expected in young Douglas-fir stand at Wind River
Experimental Forest. USDA Forest Service PNW
Research Paper PNW-RP-547, Pacific Northwest
Experiment Station, Portland, OR. 20 p.
Debrecezy, Z, and I Racz. 1995. New species and varieties of
conifers from Mexico. Phytologia 78(4):217–243.
Decker, M. 1965. Die Douglasie in Luxemburg. Paper
submit- ted in partial fulfillment of the requirements
for admission into the Luxemburg State Forestry
Commission.
Delvaux, J. 1964. Plantation de Douglas. Reprise et accrois-
sement juvénile de semis repiques en pots. Bulletin de la
Société Royale de Botanique de Belgique 71(3):141–154.
Defant, F. 1951. Local winds in Compendium of
Meteorology, American Meteorological Society, Boston,
MA, pp. 655– 670.
Del Rio, E, and AB Berg. 1979. Growth of Douglas-fir
reproduc- tion in the shade of a managed forest.
Research Paper 40, Forest Research Laboratory, Oregon
State University, Corvallis. 14 p.
Demars, EJ. 1964. Predicting Insect-Caused Damage to Douglas-
fir Seed From Samples of Young Cones. USDA. Forest
Service Res Note PSW40. Pacific Southwest Forest and
Range Experment Station, Berkeley CA. 7 p.
Dennis, FG. 1994. Dormancy— what we know (and don’t
know). HortScience 29, 1249–1255.
Dennis, FG, and LJ Edgerton. 1961. The relationship
between an inhibitor and rest in peach flower buds.
Journal of the American Society for Horticultural
Science 77:107–116.
De-Vescovi, MA, and O Sziklai. 1975. Comparative
karyotype
analysis of Douglas-fir. Silvae Genetica 24(2/3):68–72.
Devillez, F. 1970. Recherches écologiques sur la structure de
la semence et l’étalement de la germination chez
Pseudotsuga menziesii (Mirb.) Franco var. menziesii.
Comptes rendus de l'Académie des Sciences, Paris T271
Series D:2323–25.
Devillez, F. 1973. The effects of H 2O on germinative processes
in Pseudotsuga menziesii var. menziesii. Bulletin de la
Société Royale de Botaníque de Belgique 105(1):67–82.
Devitt, B. 1960. Fertilization of two improved seed-production
areas. British Columbia Forest Service Research Review
for year ended 1960. pp. 53–54.
Devlin, PF, Halliday, KJ, and GC Whitelam. 1996. The phy-
tochrome family and their role in the regulation of seed
germination, pp. 159–171 in Basic and Applied Aspects
of Seed Biology, ed. RH Ellis, M Black, AJ Murdoch,
and TD Hong. Kluwer Academic Publishers, Dordrecht.
Dewey, JE. 1970. Damage to Douglas-fir cones by
Choristoneura occidentalis. Journal of Economic
Entomology 63:804–1806.
Dexheimer J, and JC Pargney. 1991. Comparative anatomy
of the host-fungus interface in mycorrhizas. Experientia
47(4):312–321.
Deyoe, D. 1982. Soil compaction and plant physiology
Engineering, Corvallis, OR. Doidge, D. 1990. Armillaria disease and Douglas-fir beetle,
Deyoe, DR, and JB Zaerr. 1976. An improved method p. 39–41 in Pest Management Progress Vol. IX, ed. J Muir.
for ex- traction of indole-3–acetic acid from shoots 2 p. British Columbia Ministry of Forests.
of Douglas-fir. Canadian Journal of Forest
Research 6(3):429–435.
Deyoe, DR and JB Zaerr. 1976. Indole-3–acetic acid in
Douglas- fir. Plant Physiology 58:299–303.
Diaz-Vaz, OJE, and F Ojeda. 1980. Densidad
intraincremental de Pseudotsuga menziesii. I.
Variaciónes en un análisis fustal. Bosque 3(2):86–
95.
Dice, SF. 1970. The Biomass and Nutrient Flux in a
Second Growth Douglas-fir Ecosystem (A Study
in Quantitative Ecology). Doctoral dissertation,
University of Washing- ton, Seattle.
Dick, J, JM Finnis, LO Hunt, and NB Kverno. 1958.
Treatment of Douglas-fir seed to reduce loss to
rodents. Journal of Forestry 56(9):660–661.
Dick, J. 1955. Studies of Douglas-fir seed flight in
southwestern Washington. Weyerhaeuser Timber
Company, Forestry Research Note 12. Tacoma,
WA. 4 p.
Dietz, S. 1885. A külföldi fak meghonositásának
kérdéséhez (Remarks about the question of
acclimatization of exotic species). Erdészeti Lapok
24:1057–1076.
Diez, C, and A Bürgi. 1991. Wuchsleistung und
Qualitaet von Douglasie (Pseudotsuga menziesii
(Mirb.) Franco, Riesen–Lebensbaum (Thuja plicata
Donn) und Roteiche (Qeurcus rubra L.) in der
Schweiz. Berichte der Eidgenoess- ischen
Forschungsanstalt für Wald, Schnee und
Landschaft No. 329:5–46 (with Engl. summary:
Growth production and quality of Douglas-fir
(Pseudotsuga menziesii (Mirb.) Franco, western red
cedar (Thuja plicata Donn) and red oak (Quercus
rubra L.) in Switzerland).
Dighton, D. 1991. Acquisition of nutrients from
organic re- sources by mycorrhizal autotrophic
plants. Experientia 47(4):362–369.
Dimock, EJ, II. 1957. A comparison of two rodent
repellents in broadcast seeding Douglas-fir.
Research Paper 20. PNW Forest Research
Experimental Site. USDA Forest Service, Portland,
OR. 17 p.
Diniz, DMAM. 1969. Estudo do crescimento da
Pseudotsuga men- ziesii (Mirb.) Franco no norte de
Portugal. Instituto Superior de Agronomia,
Universidade Técnica de Lisboa. 70 p.
Dittmar, O. 1954. Die bisherigen Ergebnisse des
Dougla- sien–Provenienzversuches in den
Lehrreviere der Forstwirtschaftlichen Fakultaet
Eberswalde. Teil II. Die Entwicklung des
Douglasienprovenienzversuchs aus dem Jahr
1930 in Freienwalde, Abteilung 171. Archiv für
Forstwesen 3(5/6):399–431.
Dittmar, O, and E Knapp. 1967. 30 Jahre
Douglasienprove- nienzversuch Freienwalde 171.
Archiv für Forstwesen 16(6/9):847–852.
Dittmar, O, E Knapp, and B Schulsen. 1985. Ergebnisse
des internationalen Douglasienprovenienzversuches
1961 im Pleistozaen der DDR. Beiträge für die
Forstwirtschaft 19(1):8–18.
Dittmer, HJ. 1937. A quantitative study of the roots and
root harms of a winter rye plant (secale cereale).
American Journal of Botany 24:417–419.
Dittmer, HJ. 1938. A comparative study of the
subterranean members of three field grasses.
Science 88(2290) p. 482.
Dobbs, RC, DGW Edwards, J Konishi, and D Wallinger.
1974. Guideline to collecting cones of British Columbia
conifers. Brit- ish Columbia Forest Service, Canadian
Forestry Service, Joint Report 3. Victoria, BC. 98 p.
Doerksen, AH, and KK Ching. 1972. Karyotypes in the genus
Pseudotsuga. Forest Science 18(1):66–69.
References 299
Doyle, J. 1926. The ovule of Larix and Pseudotsuga. Proceedings of
Dombrosky, SA, and TD Schowalter. 1988. Inventory the Royal Irish Academy, Sect. B. 37:170–180.
monitor- ing for estimation of impact of insects on seed Doyle, J. 1945. Developmental lines in pollination mechanism
production in a Douglas-fir seed orchard in western
Oregon. Journal of Economic Entomology 81:281–85.
Domec, J-C, FC Meinzer, BL Gartner, and D Woodruff.
2006. Transpiration-induced axiel and radial tension
gradients in trunks of Douglas-fir trees. Tree Physiology
26:275–284.
Dominik, T. 1963. Mikotrofizm daglezji (Pseudotsuga
taxifolia Britton) w roznych drzewostanach w Polsce
(Occurrence of Douglas-fir (Pseudotsuga taxifolia
Britton) in various Polish stands). Prace Instytutu
Badawczego Lesnictwa No. 258:29–59. (Translation
published by US Department of Agriculture and the
National Science Foundation, Washington, DC, by the
Scientific Publications Foreign Cooperation Center of
the Central Institute for Scientific, Technical and
Economic Information, Warsaw, Poland, 1966. TT 65–
50353.
Domisch, T, L Finer, and T Lehto. 2000. Soil temperature
effects on biomass and carbohydrates. Tree Physiology
21:465–471.
Dong, PH. 1970. Wuchsleistung and biologisch–waldbauliches
Verhalten der Douglasie in Kulturversuchen der
Niedersäch- sischen Forstlichen Versuchsanstalt. Doctoral
dissertation, Forstliche Fakultät, GeorgAugust
Universität, Göttingen.
Dong, PH. 1973. Anbaufolgerungen aus den bisherigen
Ergeb- nissen über Ausfälle bei verschiedenen
Douglasienher- künften im Aufwuchsstadium.
Allgemeine Forstzeitung 28(48):1056–1058.
Doorenbos, J. 1953. Review of the literature of dormancy in
buds of woody plants. Medelingen van de Landbouwhoge-
school Wageningen/Nodertaad 53(1):1–24.
Dornbos, DL Jr. 1995. Seed vigor, pp. 45–80 in Seed
Quality: Basic Mechanisms and Agricultural
Implications, ed. AS Basra. Food Products Press
(Hawworth Press, Inc.), New York. 389 p.
Dosskey MG, RG Linderman, and L Boersma. 1990.
Carbon- sink stimulation of photosynthesis in
Douglas-fir seed- lings by some ectomycorrhizae.
New Phytologist 115, 269–274.
Dosskey, MG, L Boersma, and RG Linderman. 1991. Role
for the photosynthate demand of ectomycorrhizas in
the response of Douglas-fir seedlings to drying soil.
New Phytologist 117:327–334.
Dosskey, MG, L Boersma, and RG Linderman. 1993.
Effect of phosphorus fertilization on water stress in
Douglas-fir seedlings during soil drying. Plant and
Soil 150:33–39.
Doumas, P, and JB Zaerr. 1988. Seasonal changes in levels
of cyto-kinin-like compounds from Douglas-fir
extrolate. Tree Physiology 4:1–8.
Doumas, P, JW Morris, E Chien, M Bonnet-Masimbert, and
JB Zaerr. 1986. A possible relationship between a cyto-
kinen conjugate and flowering in Douglas-fir, pp. 288–
96 in Ninth North American Forest Biology Workshop,
Physi- ologic and genetic basis of tree decline, ed. CG
Taver and TE Hennessey. Oklahoma State University,
Stillwater. 321 p.
Doumas, P, M Bennet-Masimbert, and JB Zaerr. 1989.
Evidence of Cytokinen bases, ribosides and glucosides
in roots of Douglas-fir, Pseudotsuga menzeisii. Tree
Physiology 5:63–72.
Dowding, P. 1987. Wind pollination mechanisms and
aerobiol- ogy. International Review of Cytology
107:421–437.
Downie, B, U Bergsten, BSP Wang, and JD Bewley. 1993.
Conifer seed germination is faster after membrane
tube invigoration than after prechilling or osmotic
priming. Seed Science Research 3:259–270.
Doyle, J. 1918. Observations on the morphology of Larix lep-
tolepsis. Scientific Proceedings Royal Dublin Society
15:310– 327+2 pls.
outplanted seedlings and perfor- mance of the
in the Coniferales. Scientific Proceedings of the Royal Dublin pathogen. New Forests 7:143–149.
Society 24:43–62. Dumroese, RK, RL James, DL Wenny. 1995. Interactions be-
Doyle, J, and M O’Leary. 1935. Pollination in Tsuga, tween copper coated containers and fusarium root disease.
Cedrus, Pseudotsuga, and Larix. Scientific
Proceedings of the Royal Dublin Society 21:191–204.
Drew, AP, and WK Ferrell. 1979. Seasonal changes in
the water balance of Douglas-fir (Pseudotsuga
menziesii) seedlings grown under different light
intensities. Candian Journal of Botany 87:666–674.
Drew, JT. 1957. Relationship of Locality and Rate of
Growth to Density and Strength of Douglas-fir.
Report No. 2078 (In- formation reviewed and
reaffirmed 1965). USDA, Forest Service, Forest
Products Laboratory, Madison, WI.
Droppelmann, FJ. 1986. Evaluacion de un ensayo de
proceden- cias de pino oregón (Pseudotsuga
menziesii (Mirb.) Franco) de 17 años de edad
(Fundo Las Palmas, Valdivia). Ingeniero Forestal
thesis, Universidad Austral de Chile, Facultad de
Ciencias Forestales, Valdivia.
Ducci, F, and A Tocci. 1987. Primi resultati della
sperimen- tazione IUFRO 1969–1970 su Pseudotsuga
menziesii (Mirb.) Franco nell’ Appennino centro–
settentrionale (with Eng- lish summary: First results
of IUFRO 1969–70 experimen- tation on
Pseudotsuga menziesii (Mirb.) Franco in northern
and central Apennines.) Annali dell’ Istituto
Sperimentale per la Selvicoltura 18:215–289.
Ducic T, L Leinemann, R Finkeldey, and A Polle. 2006.
Up- take and translocation of manganese in
seedlings of two races of Douglas-fir (Pseudotsuga
menziesii var. viridis and glauca). New Phytologist
170:11–20.
Dudarev, AD, VI Dusha, NG Kosarev, and VV
Uspenski. 1975. Growth and yield of
representatives of North American tree species on
the Black Sea coast of the Caucasus. Lesnoi
Zhurnal 18(2):28–32.
Duddridge, JA, A Malibori, and DH Read. 1980.
Structure and function of mycorrhizal
rhizomorphs with special refer- ence to their role
in water transport. Nature 287:834–836.
Dudetskaya, EM, and AV Lukin. 1977. (Pseudotsuga
menziesii in the central chernozem zone of the
RSFSR) Lzhetsuga menzieza v tsentral’no–
chernozemnykh oblastyakh RS- FSR. Byulleten’
Glavnogo Botanicheskogo Sada No. 106:22–25.
Original not seen, cited from Forestry Abstracts
40:1021, 1979.
Duffield, JW. 1950. Techniques and possibilities for
Douglas- fir breeding. Journal of Forestry
48(1):41–45.
Duffield, JW. 1956. Damage to western Washington
forests from November 1955 cold wave. USDA
Forest Service. PNW Forest and Range Experiment
Station Research Note 129, Portland, OR. 5 p.
Dumont-BéBoux N, and P von Aderkas. 1997. In vitro
pollen tube growth in Douglas-fir. Canadian
Journal of Forest Research 27:674–678.
Dumont-BéBoux, N, B Anholt, and P von Aderkas.
1999. In vitro Douglas fir pollen germination:
influence of hydra- tion, sucrose and polyethylene
glycol. Annals of Forest Science 56(1):11–18.
Dumont-BéBoux, N, M Weber, Y Ma, and P. von
Aderkas. 1998. Intergeneric pollen–
megagametophyte relationships of conifers in vitro.
TAG Theoretical and Applied Genetics 97 (5):881–
887.
Dumroese, RK, and RJ James. 2005. Root diseases in
bare root and container nurseries of the Pacific
Northwest: epidemiology, management, and
effects on outplanting performance. New Forests
30:185–202.
Dumroese, RK, RL James, and DL Wenny. 1993.
Fusarium root infection of container-grown
Douglas-fir: effect on survival and growth of
300 Douglas-fir: The Genus Pseudotsuga
A Preliminary Report. OSDA. Forest Service Northwest For- est and Range Experiment Station, Portland,
Northern Region Insect and Disease Report 95–9. OR. 130 p.
Dunberg, A, and PC Oden. 1983. Gibberellins and conifers,
pp. 221–295 in The Biochemistry and Physiology of
Gibberellins. Vol 2, ed. A Crozier. Praeger, New York.
Dunlap, LH. 1964. Differentiation of Coastal and Interior Prov-
enances Using Morphology of Seed. BSF Thesis,
University British Columbia, Faculty of Forestry,
Vancouver.
Dunsworth, GB. 1997. Plant quality assessment: an industrial
perspective. New Forests 13:439–445.
Durley, RC, RP Pharis, and JAD Zeevaart. 1975. Metabo-
lism of [H3] gibberellin A 2O by plants of Bryophyllum
diagnemontanum under long and short-day conditions.
Plants 126:139–149.
Duryea ML, and SK Omi. 1987. Top-pruning Douglas-fir
seed- lings: morphology, physiology and field
performance. Canadian Journal of Forest Research
17:1371–1378.
Duryea, ML, and DP Lavender. 1982. Water relations,
growth, and survival of root-wrenched Douglas-fir
seedlings. Canadian Journal of Forest Research
12(3):545–555.
Durzan, DJ, and RA Campbell. 1979. Red light inhibits
female cone production while spruce trees, pp. 57–62 in
Proceed- ings: Flower Symposium, ed. F Bonner. 375 p.
Dusha, VI. 1977. Povysit’ esteticheskuyu i
bal’neologicheskuyu tsennost’lesov (Improving the
aesthetic and balneological value of forests). Lesnoe
Khozyaistvo 1977, No. 10:73–74. Original not seen,
cited from Forestry Abstracts 39:4038, 1978.
Ebell, LF. 1962. Growth and Cone Production Responses of
Doug- las-fir to Chemical Fertilization. Report on
Project BC 1, Canada Department of Forestry, BC
District, Mimeo. Victoria, BC.
Ebell, LF. 1967. Cone production induced by drought in pot-
ted Douglas fir. Canada Department of Forestry
Bimonthly Research Notes 23:26–27.
Ebell, LF. 1970. Physiology and biochemistry of flowering
of Douglas-fir in sexual reproduction of forest trees.
Pro- ceedings IUFRO Section 22 working group, 1970.
Finnish Forest Research Institute, Helsinki, Varparanta,
Finland.
Ebell, LF. 1971. Girdling: its effect of carbohydrate status
and on reproductive bud and cone development of
Douglas- fir. Canadian Journal of Botany 49:453–466.
Ebell, LF. 1972. Cone-production and stem-growth response
of Douglas-fir to rate and frequency of nitrogen
fertilization. Canadian Journal of Forest Research
2(3):327–338.
Ebell, LF, and EE McMullan. 1970. Nitrogenous substances
associated with differential cone production responses
of Douglas-fir to ammonium and nitrate fertilization.
Canadian Journal of Botany 48(12):2169–2177.
Ebell, LF, and RL Schmidt. 1964. Meteorological Factors
Affecting Conifer Pollen Dispersal on Vancouver Island.
Department of Forestry Publication 1036. Government of
Canada, Pacific Forest Research Centre, Victoria, BC.
Edgar, MJ, D Lee, and BP Quinn. 1992. New Zealand Forest
Industries Strategy Study. New Zealand Forest Industries
Council Publication.
Edgren, JW. 1970. Growth of frost-damaged Douglas-fir
seedlings. Research Note PNW-121, USDA Forest Service
Pacific Northwest Research Station, Portland, OR. 8 p.
Edgren, JW. 1968. Potential damage to forest tree seed
During Processing, Protective Treatment and Dissemination.
Research Note PNW-89.USDA Forest Service Pacific
Northwest Research Station, Portland, OR.
Edgren, JW, and DM Trappe. 1970. Growth of Douglas-fir,
ponderosa pine, and western larch seedlings following
seed treatment with 30 percent hydrogen peroxide.
Research Note PNW 121, USDA Forest Service Pacific
Edwards, DGW. 1980. Maturity and quality of tree container nursery production, pp. 34–42 in Dormancy
seeds–a state of the art review. Seed Science and and barriers in germination. Proceedings of an
Technology 8:625– 657. International Symposium of IUFRO project group. pp.
Edwards, DGW. 1985. Cone prediction, collection, and 204–00 (Seed Problems) ed. DGW Edwards. Forestry
pro- cessing, pp. 78–102 Proceedings Coniferious Canada, Pacific Forestry Centre, Victoria, BC.
Tree Seed Inter- mountain West Symposium, comp. El-Lakany, MH, and O Sziklai. 1971. Intraspecific variation in
RC Schearer. General Technical Report INT-203.
USDA Forest Service Inter- mountain Research
Station, Ogden, UT. 289 p.
Edwards, DGW. 1986. Special rechilling techniques for tree
seeds. Journal of Seed Technology 10(2):151–171.
Edwards, DGW, and YA El-Kassaby. 1996. The biology
and management of coniferous forest seeds: genetic
perspec- tives. Forestry Chronicles 72:481–484
Edwards, DGW, and PE Olsen. 1972. R-55 Rodent
repellent: effect on germination in Douglas-fir and
western hemlock. Canadian Journal of Forest
Research 2:256–263.
Edwards, DWG, YA El-Kassaby, and DP Lavender.
Unpub- lished ms. Forest Tree Seeds in the Pacific
Northwest: Biology Collection and Post Harvest
Handling. 63 p.
Edwards, MV. 1956. The design layout, and control of
prov- enance experiments. Silvae Genetica 5:169–
181.
Edwards, MV. 1957. Pseudotsuga Carrière, pp. 118–126 in
Exotic Forest Trees in Great Britain, ed. J
Macdonald, RF Wood, MV Edwards, and JR
Aldhous. Forestry Commission Bulletin No. 30,
HMSO.
Eis, S. 1970. Reproduction and reproductive irregularities
of Abies Iasiocarpa and A. grandis. Canadian Journal
of Botany 48:141–143.
Eis, S. 1972. Climatic prerequisites for Douglas-fir and
grand fir cone crops, p. 43 in North American
Forest Biology Work- shop, Programme Abstracts,
2nd ed., comp. DP Lavender.
Eis, S. 1973a. Cone production of Douglas-fir and
grand-fir and its climatic requirements. Canadian
Journal of Forest Research 3(1):61–70.
Eis, S. 1973b. Root system morphology of western
hemlock, western redcedar, and Douglas-fir.
Canadian Journal of Forest Research 4:28–38.
Eis, S, and D Craigdellie. 1981. Reproduction of
conifers. Cana- dian Forestry Service / Pacific
Forest Research Center. Victoria, BC Canada. BC–
X219.
Eis, S, EH Garman, and LF Ebell. 1965. Relation
between cone production and diameter increment of
Douglas-fir (Pseudotsuga menziesii (Mirb.)
Franco), grand fir (Abies grandis (Dougl.) Lindl.)
and Western white pine (Pinus monticola Dougl).
Canadian Journal of Botany 43:1553–1559.
El-Kassaby, YA, and Park, YS. 1990. Harvest index and
wood density in a Douglas-fir early progeny test. In
Proceed- ings, Joint Meeting, WFGA and IUPRO
Working Parties 2.02–05,06,12,14. Section 4,45,
Aug. 1990. Olympia, WA. 11 p.
El-Kassaby, YA, and O Sziklai. 1982. Genetic variation
of al- lozyme and quantitative traits in a selected
Douglas-fir (Pseudotsuga menziesii (Mirb.) Franco)
population. Forest Ecology and Management
4(2):115–126.
El-Kassaby, YA, Edwards, DGW, and C Cook. 1990.
Impact of crop management Practices on Seed
Yield in a Douglas-fir Seed Orchard. Silvae
Genetica 39 5–6:226–230.
El-Kassaby, YA, DGW Edwards, and DW Taylor. 1992.
Genetic control of germination parameters in
Douglas-fir and its importance for domestication.
Silvae Genetica 41:48–54.
El-Kassaby, YA, K Chaisurisri, DGW Edwards, and
DW Tay- lor. 1993. Genetic control of germination
parameters of Douglas-fir, Sitka spruce, western
redcedar, and yellow- cedar and its impact on
References 301

nuclear characteristics of Douglas-fir. Advancing ed. J-D Viémont and J Crabbé. CAB International, Paris.
Frontiers of Plant Science 28:363–378. Eriksson, A, and M Widerlund. 1992. Douglasgran–
El-Lakany, MH, and O Sziklai. 1973. Further investigations revidering av provytor. Examensarbete no. 6. Sveriges
of intraspecific variation in DNA contents of Douglas-fir Lantbruksuni- versitet, Skogsmästarskolan 1992.
(Pseudotsuga menziesii (Mirb. Franco). Egyptian Journal Erteld, W. 1948. Bisherige Ergebnisse eines Douglas-
of Genetics and Cytology 2(2) 345–354. Provenienz- versuches im Forstamt Freienwalde.
Elwes, HJ, and A Henry. 1909. The trees of Great Britain Forstwirtschaft Hol- zwirtschaft 2(9):135–138.
and Ireland, Vol. 4. Privately printed, Edinburgh. Etverk, I. 1978. Observations on cultivating some foreign tree
Emmingham, WH. 1977. Comparison of selected Douglas-fir species in Estonia, pp. 369–380 in Vol. 1: Background pa-
seed sources for cambial and leader growth patterns in pers and Douglas-fir provenances. Proceedings IUFRO
four western Oregon environments. Canadian Journal of joint meeting of Working Parties S2.02–05 Douglas-fir
Forest Research 7(1):154–164. provenances, S2.02–06 Lodgepole pine provenances,
Emmingham, B. 1997. Thinning Westside forests and S2.02–12 Sitka spruce provenances, S2.02–14 Abies
stand development, density relationships and provenances. Vancouver, Canada, 1978. British Columbia
management objectives. In Symposium on Thinning Ministry of Forests, Information Services Branch, Victoria,
in Westside For- ests. Adaptive COPE Program, SAF, BC.
College of Forestry, Oregon State University, Evans, LT. 1971. Flower induction and the florigen concept.
Corvallis. Review of Plant Physiology 22:365–394
Emmingham, WH, and RH Waring. 1973. Conifer growth Evelyn, J. 1664. Sylva: or a Discourse of Forest Trees & the
under different light environments in the Siskiyou Propa- gation of Timber, Vol. II. Reprinted in London by
Mountains of southwestern Oregon. Northwest Arthur Doubleday & Co., Ltd.
Science 47(2):89–99. Everett, RL, R Schellhaas, D Keenum, D Spurbeck, and P
Emmingham, WH, and RH Waring. 1977. An index of photo- Olson. 2000. Fire history in the ponderosa
synthesis for comparing forest sites in western Oregon. pine/Douglas- fir forests on the east slope of the
Canadian Journal of Forest Research 7(1):165–174. Washington Cascades. Forest Ecology and
Enescu, V. 1979. Cercetari de provenienta la duglas si larice. Management 129:207–225.
Minis- terul Economiei Forestiere si Materialelor de Eyre, FH, ed. 1980. Forest Cover Types of the United States
Constructuii, Departamentul silviculturii, Institutul de and Canada. Society of American Foresters,
Cercetari se Amenajari Silvice, Bucharest. Washington, DC. 148 p.
Enescu, V. 1984. Cercetari de provenienta la duglas si Fabricius, O. 1926. Douglas– og Sitkagran. Dansk
larice. Ministerul Silviculturii, Institutul de Cercetari si Skogeforenings Tidskrift 11:405–541.
Ame- najari Silvice, Seria a 11–a. Fashler, AMK, YA El–Kassaby, and O Sziklai. 1987.
Enescu, V. 1987. Climate and the choice of seed orchard Interprov- enance variation in the IUFRO Douglas-fir
sites. provenance/ progeny trial, pp. 187–204 in Proceedings
Forest Ecology and Management. 19:257–265. of IUFRO Work- ing Party S2.02–05 on breeding strategy
Entry, JA, SK Hagle, and K Cromack. 1990. The effect of for Douglas-fir as an introduced species. Vienna, Austria
Armil- laria attack on the nutrient status of Inland June 10–14, 1985. FBVA Berichte No.21. Schriftenreihe
Douglas-fir. European Journal of Forest Pathology Forstliche Bundersver- suchsanstalt Vienna, Austria.
20:269–274. Fatzinger, CW, and WC Asher. 1971. Mating behavior and
Entry, JA, KJ Cromack Jr, RG Kelsey, and NE Martin. evidence for a sex hormone of Dioryctria abielella
1991a. Response of Douglas-fir to infection by (Lepidop- tera: Pyralidae (Phyceitlinae)). Annals of the
Armillaria ostoyae after thinning or thinning plus Entomological Society of America 64(3):612–620.
fertilization. Phytopathol- ogy 81:682–689. Faulkner, R. 1966. A review of flower induction experiments
Entry, JA, K Cromack Jr., RG Kelsey, and NE Martin. and trials 1948-63. UK Forestry Commission, Report on
1991b. Response of western coniferous seedlings to Forest Research for the year ended March 1965, pp. 207–
infection by Armillaria ostoyae under limited light and 219.
nitrogen. Phytopathology 81(1):89–94. Fayle, DCF. 1968. Radial growth in tree roots. Technical
Entry, JA, PK Donnelly and K Cromack, Jr. 1992a. The Report No. 9, Faculty of Forestry, University of Toronto.
influ- ence of carbon nutrition on Armillaria ostoyae 183 p.
growth and phenolic degradation. European Journal of Fayle, DCF. 1980. Secondary thickening in tree roots and
Forest Pathology 22(3):149–156. environmental influences, pp. 93-118 in Environment
Entry, JA, NE Martin, RG Kelsey, and K Cromack, Jr. and Root Behavior, ed. DN Sen. Geobios
1992b. Chemical constituents in root bark of five species International, Jodhpur, India.
of western conifer sapling and infection by Armillaria Fechner, GH. 1979. The biology of flowering and fertilization,
ostoyae. Phytopathology 82, 393–397. pp. 1–24 in Proceedings IUFRO Meeting on Flowering
Entry, JA, PK Donnelly, and K Cromack, Jr. 1993. Effect of and Seed Development in Trees, May 15–18, 1978,
nitrogen and carbon sources on lignin and cellulose deg- Starkville, Mississippi, USA, ed. F. Bonner.
radation by Armillaria ostoyae. European Journal of Fedorov, EA. 1981. The introduction of Pseudotsuga in the
Forest Pathology 23(3):149–156. Kalingrad region. Lesnoe Khozyaistvo 1981, No. 1:34–
Erdmann, GG. 1966. Promising Conifers for Western Iowa. 35. Original not seen, cited from Forestry Abstracts
Re- search Paper NC-8. USDA Forest Service, North 43:4065, 1982.
Central Forest Experiment Station, St. Paul, MN. Fenton, RT. 1967. The role of Douglas-fir in Australasian
Erdtman, G. 1943. An introduction to pollen analysis. Chronica for- estry. New Zealand Journal of Forestry 12(1):4–41.
Botanica Company, Waltham, MA. Fenton, R. 1978. Risks or the case for diversification of
Eremko, RD, DGW Edwards, and D Wallinger. 1989. A produc- tion plantation species and the role of Douglas
Guide to Collecting Cones of British Columbia Conifers. fir, pp. 104–106, 433–446 in A review of Douglas fir in
Joint Publi- cation of Forestry Canada and British New Zealand, 16-19 September, 1974, ed. RN James and
Columbia Ministry of Forests, Victoria, BC. 114 p. EH Bunn. New Zealand Forest Service, Forest Research
Institute, Rotorua.
Erez, A. 2000. Bud dormancy: a suggestion for the control
mechanism and its evolution, pp. 23–33 in Dormancy Fernandez, DJA. 1964. Crecimiento de un bosque de pino
in Plants: From Whole Plant Behaviour to Cellular oregón, Pseudotsuga menziesii, en San Martin de los
Control, Andes, Neuquén, Argentina. Revista Forestal Argentina
8(2):37–41.
Fernando, DD, JN Owens, P von Aderkas, and T Takaso. 1997.
302 Douglas-fir: The Genus Pseudotsuga
In vitro pollen tube growth and penetration of female Fletcher, AM, and Barner, H, 1978. The procurement of seed
gametophyte in Douglas-fir (Pseudotsuga menziesii). for provenance research with particular reference to
Sexual Plant Reproduction 10:209–216. collections in NW America, pp. 141–154 in Proceedings
Fernando DD, JN Owens, P von Aderkas. 1998. In vitro of the IUFRO Joint Meeting Working Parties,
fertil- ization from co-cultured pollen tubes and female Vancouver, BC, Canada, Vol. 1, ed K. Ching and O.
game- tophytes of Douglas fir (Pseudotsuga menziesii). Sziklai. Ministry of Forests, BC.
Theoretical and Applied Genetics 96:1057–1063 Fletcher, AM, and CJA Samuel. 1990. Growth and branch-
Ferraris, P. 1993. Adattamento e accrescimento di diverse ing characteristics in the IUFRO origins of Douglas-fir
provinienza di douglasia del nord America: 16 years after planting in Britain, Section 2.71:1–11 in
osservazioni in un implanto comparativo de 22 anni. Proceedings of the Joint Meeting of Western Forest
Instituto Speri- mentali Selvicultura. [Adaptation and Genetics Assoc. and IUFRO Working Parties S2.02–05, –
growth of different North American provenances of 06, –12, –14.
Douglas-fir: Observation on a 22 year old comparative Olympia, Washington, August 20–24, 1990.
plantation]. Ann. Inst. Spec. Selv. 24:63-78. Weyerhaeuser Co., Centralia, WA.
Ferré, Y, and H Gaussen. 1945. Le rameau phylétique Pinus, Flöhr, W. 1954. Die bisherigen Ergebnisse der Doug-
Pseudolarix, Keteleeria. Traveaux du Laboratoire Foristier de lasien–Provenienzversuche in den Lehrrevieren der
Toulouse, Tome I, Vol. 4, art. VIII (Original not seen, Forstwirtschaftlichen Fakultät Eberswalde. Teil I. Die
cited from van Campo–Duplan, 1950). Entwicklung des Douglasien–Provenienzversuches
Ferrell, WK, and ES Woodard. 1966. Effect of seed origin on aus dem Jahr 1910 in Chorin, Abteilung 90e. Archiv
drought resistance of Douglas-fir (Pseudotsuga für Forst- wesen 3(5/6):385–399.
menziesii (Mirb.) Franco.) Ecology 47:499–503. Flöhr, W. 1958. Kennzeichnung, Varietäten und Verbreitung
Fielder, P, and JN Owens. 1989. A comparative study of der Douglasie, pp. 4–10 in Die Douglasie und ihr Holz.
shoot and root development of interior and coastal K. Göhre ed. Akademie Verlag, Berlin.
Douglas-fir seedlings. Canadian Journal of Forest Florin, R. 1954. The female reproductive organs of conifers
Research 19:539–549. and taxads. Biological Reviews of the Cambridge
Filip, GM, and LM Ganio. 2004. Early thinning in mixed- Philosophical Society 29:367–389.
species plantation of Douglas-fir, hemlock, and true fir Florin, R. 1963. The distribution of conifer and taxad genera
affected by Armillaria root disease in west central in time and space. Acta Horti Bergiani 20(4):121–312.
Oregon and Washington: 20 year results. Western Flous, F. 1936. Classification et évolution d’un group d’Abié-
Journal of Applied Forestry 19(1):25–33. tinées. Traveaux du Laboratoire Forestier de Toulouse,
Finlay, RD. 2008. Ecological aspects of mycorrhizal symbio- Tome I, Vol. 2, art. 17.
sis: with special emphasis on the functional diversity of Foerst, K. 1980. Standort, Wuchsleistung und Ernaeh-
interactions involving the extraradical mycelium. rungszustand älterer bayerischer Bestände der grünen
Journal of Experimental Botany 59(5):1115–1126. Douglasie (Pseudotsuga menziesii [Mirb.] Franco) var.
Finnis, JM. 1950. Seed Maturity in Douglas-fir. Research Note menziesii). Mitteilungen der Staatsforstverwaltung Bayerns
#18, BC Forest Service. 8 p. No. 41.
Finnis, JM. 1953. A note on the bud count method for Fogel, R. 1983. Root turnover and productivity of coniferous
fore- casting cone crops of Douglas-fir. Forestry forests. Plant and Soil 71:75–85.
Chronicle 29(2):22–27. Fogel, R. 1990. Root turnover and production in forest trees.
Finnis, JM. 1955. The Use of Tetramine in the Direct Seeding HortScience 25(3):270–273.
of Douglas-fir in Coastal British Columbia. British Fogel, R, and G Hunt. 1979. Fungal and arboreal biomass in
Columbia Forest Research Note #3. Dept. of Lands and a western Oregon Douglas-fir ecosystem: distribution
Forests. Victoria, BC. 22 p. patterns and turnover. Canadian Journal of Forest
Finnis, JM. 1957. Direct Seeding of Douglas-fir with Special Research 9:245–256.
Refer- ence to Survival. Master’s thesis, Oregon State Fogal, WH, G Jobin, HO Schooley, SJ Coleman, and MS
University, Corvallis Wolynetz. 1996. Stem incorporation of gibberellins to
Fisher, JT, and JG Mexal. 1984. Nutrition management: A promote sexual development of white spruce, Norway
Physical basis for yield improvement, pp. 271–199 in spruce, and jack pine. Canadian Journal of Forest
Seedling Physiology and Reforestation Success, ed. ML Research 26(2):186–195.
Dur- yea and GN Brown. Nijhoff/Junk, Boston, MA. 326 Folk, RS, and SE Grossnickle. 1997. Determining field
p. perfor- mance potential with the use of limiting
Fitter, AH. 1991. Costs and benefits of mycorrhizas - environmental conditions. New Forests 13:121–138.
Implica- tions for functioning under natural conditions. Folk, RS, SC Grossnickle, P Axelrod, and D Trotter. 1999.
Experientia 47(4):350–355. Seed- lot, nursery, and bud dormancy effects on root
Fitzpatrick, HM. 1966. The Forests of Ireland. An Account of electrolyte leakage of Douglas-fir (Pseudotsuga
the Forests of Ireland from Early Times until the Present menziesii) seedlings. Canadian Journal of Forest
Day. Society of Irish Foresters, n.p. Research 29:1269–1281.
Flemion, F. 1938. A rapid method for determining the vi- Fombroksky, SH, and TD Schowalter. 1988. Inventory
ability of dormant seeds. Contributions Boyce Thompson Monitor- ing for estimating impact of insects on seed
Institute. 9:339–351. production in a Douglas-fir seed orchard in western
Flemion, F. 1941. Further studies on the rapid determination Oregon. Journal of Economic Entomology 81:281–285.
of the germinative capacity of seeds. Contributions Fonda, RW, and Bliss, LC. 1969. Forest vegetation of the
Boyce Thompson Institute 11(6):455–484. montane and subalpine zones, Olympic National Park,
Flemion, F. 1948. Reliability of the excised embryo method Washington. Ecological Monographs 39:279–301.
as a rapid test for determining the germinative capacity Fontes, L, M Tomé, F Thompson, A Yeomans, JS Luis, and
of dormant seeds. Contributions Boyce Thompson P Savill. 2003. Modelling the Douglas-fir (Pseudotsuga
Institute 15:229–240. menziesii (Mirb.) Franco) site index from site factors in
Flemion, F, and H Poole. 1948. Seed Viability Tests with 2, Portugal. Forestry 76(5):491–507.
3, 4-Triphenyl Tetrazolium chloride. Contributions Foote, RH. 1956. Gall midges associated with cones of
Boyce Thompson Institute 15(4):243–258. Western forest trees. (Diptera: Itonididae). Journal of the
Washington
References 303
Furniss, RL, and VM Carolin. 1977. Western Forest Insects.
Academy of Sciences 46(2):48–57. USDA Forest Service, Miscellaneous Publication 1339,
Washington, DC. 654 p.
Forward, BS, Transberger, TH, and S Misra. 2001. Charac-
terization of proteinase activity in stratified Douglas-fir
seeds. Tree Physiology 21:625–629.
Foster, AS. 1942. Practical Plant Anatomy. D. Van Nostrand
Company, Inc., New York.
Fourchy, P. 1954. Etudes sur le développement et la produc-
tion de quelques peuplements de Douglas (Pseudotsuga
douglasii Carr.) Annales de l’Ecole Nationale des Eaux
et Forêts 14(1):87–151.
Fovernied, D. 1946. Les richesses forestières des Paysbas.
Revue internationale du bois 13(106):89–92.
Fowells, HA. 1965. Silvics of Forest Trees of the United
States. Agriculture Handbook No. 271, USDA Forest
Service, Washington, DC.
Fowler, DP, and TW Dwight. 1964. Provenance differences
in the stratification requirements of white pine.
Canadian Journal of Botany 42:669–675.
Franco, JDA. 1954. On the legitimacy of the combination
Pseudotsuga menziesii (Mirb.) Franco. Boletim da
Sociedade Broteriana (sér. 2) 28:115–116.
Frank, CJ, and MJ Jenkins. 1987. Impact of the western
spruce budworm (Lepidoptera: Tortricidae) on buds,
developing cones, and seeds of Douglas-fir in West
Central Idaho. Environmental Entomology 16(1):304–
308.
Franklin, JF, and CT Dyrness. 1973. Natural vegetation of
Oregon and Washington. USDA Forest Service, General
Technical Report PNW-8. Pacific Northwest Forest and
Range Ex- periment Station, Portland, OR. 417 p.
(Reprinted 1988, Oregon State University Press,
Corvallis.)
Franklin, JF, and M Hemstrom. 1981. Aspects of
succession in the coniferous forests of the Pacific
Northwest, pp. 212–229 in Forest Succession: Concepts
and Application, ed. DC West, HH Shugart, and DB
Botkin, Springer Verlag, New York.
Franklin, JF, and RH Waring. 1980. Distinctive features of
the northwestern coniferous forest: Development,
structure, and function, pp. 59–86 in Forests: Fresh
Perspectives from Ecosystem Analysis, Proceedings of the
40th Annual Biology Colloquium, ed. RH Waring,
Oregon State University Press, Corvallis.
Fraser, I. 1978. The Douglas fir resource and current
utilization, pp. 347–354 in A review of Douglas-fir in
New Zealand. RN James, and EH Bunn eds. FRI
Symposium No. 15. Forest Service, Forest Research
Institute.
Freitas, ASB. 1989. Perimetro Florestal de Manteigas.
Direcção– Geral das Florestas, Lisbon, Portugal.
(Original not seen, cited from Luis 1989).
Friend, AL, MR Eide, and TM Hinckley. 1990. Nitrogen
stress alters root proliferation in Douglas-fir seedlings.
Canadian Journal of Forest Research 20:1524–1529.
Froehlich, HA, DH McNabb, and F Gaweda, 1982.
Average annual precipitation 1960-1980, in Southwest
Oregon. OSU Extension Service Miscellaneous
Publication 8220, Corvallis, OR. 8 p.
Frothingham, EH. 1909. Douglas fir: A study of the Pacific
Coast
and Rocky Mountain Forms. USDA Forest Service, Circular
150. Government Printing Office, Washington, DC.
38 p. Fürstenberg, M von. 1923. Die Einführung einer
winterharten
Form der Douglastanne in Deutschland. Mitteilungen
der
Deutschen Dendrologischen Gesellschaft 33:79–90.
Füerstenberg, M von. 1924. Die Einführung einer
winterharten Form der Douglastanne in Deutschland
(Nachtrag zu dem Aufsatz, Jahrbuch 1923, S. 79 ff.)
Mitteilungen der Deutschen Dendrologischen
Gesellschaft 34:372.
Gaussen, H. 1966. Les gymnospermes actuelles et fossiles.
Galoux, A. 1952. Etude des conditions de croissance Traveaux du Laboratoire Forestier de Toulouse, Tome II,
des essences forestières américaines susceptibles Vol. I, Fasicule VIII :481–672.
d’intéresser la sylviculture européenne. Geelhand, G. 1954. Le semis naturel de douglas en
Conclusions finales pour la Belgique. I. Essences Campine.
forestières américaines les plus importantes à Bulletin de la Société Royale de Botanique de Belgique
retenir pour la sylviculture belge. Revue de
l’Agriculture 5(6):591–594.
Galoux, A. 1956. Le sapin de Douglas et la
phytogéographie. Station de Recherches de
Groenedaal–Hoeillaart, Traveaux– Series B, No. 20.
Gamble, WE, WT Adams, and J Tappeiner II. 1996.
Survival and Growth of Douglas-fir Seed Sources in
the Hospital Tract Range Wide Source Archive
Plantation. Forest Research Laboratory Research
Contribution 12, Oregon State Uni- versity,
Corvallis.
Ganghofer, Av. 1884. Arbeitsplan für die
Anbauversuche mit ausländischen Holzarten.
Festgestellt bei der Berathung zu Braunschweig im
August 1881 pp. 169–190 in Volume 2, Das
forstliche Versuchswesen, Augsburg. (Published also
in Danckelmann and Mundt, Jahrbuch der
Preussischen Forst-und Jagdgesetzgebung, Vol. 14,
1882).
Garman, EH. 1951. Seed Production by Conifers in the
Coastal Region of British Columbia Related to
Dissemination and Regeneration. British Columbia
Forest Service Technical Publication T35, Research
Division, Department of Lands and Forests,
Victoria, BC. 47 p.
Garman, EH. 1955. Regeneration Problems and Their
Silvicultural Significance in the Coastal Forests of
British Columbia. Brit- ish Columbia Forest Service,
Technical Publication T41, Research Division,
Department of Lands and Forests, Victoria, BC. 67 p.
Garman, EH, and AL Orr-Ewing. 1949. Direct-seeding
Experi- ments in the Southern Coastal Region of
British Columbia, 1923–1949. British Columbia
Forest Service, Technical Publication Technical
Publication 31, Research Division, Department of
Lands and Forests, Victoria, BC. 46 p.
Gashwiler, JS. 1967. Conifer seed survival in a Western Oregon
clearcut. Ecology 48(3):431–438.
Gashwiler, JS. 1969. Seed fall of three conifers in West-Central
Oregon. Forest Science 15(3):290–295.
Gashwiler, JS, AL Ward. 1968. Oregon Junco foods in
conifer- ous forests. The Murrelet 49(3):29–36.
Gast, WR, DW Scott, C Schmitt, D Clemens, S Howes,
CG Johnson, R Mason, F Mohr, and RA Clapp.
1991. Blue Mountains Forest Health Report: New
Perspectives In Forest Health. USDA Forest Service
Pacific Northwest Region, Malheur, Umatilla and
the Wallowa-Whitman National Forests of NE
Oregon and SE Washington.
Gathy, P. 1956. Aperçu des recherches en matière de
géné- tique forestière. Bulletin de la Société Royale
de Botanique de Belgique 63(10):393–433.
Gathy, P. 1957. Recherches belges sur la variabilité
génétique des espèces forestières. Silvae Genetica
6(1/2):32–38.
Gathy, P. 1961. Rapport préliminaire sur un test
d’origines de douglas vert (Pseudotsuga taxifolia
Britt.) Travaux–Série B, No.26. Station de
Recherches des Eaux et Forêts, Groe- nendaal–
Hoeilaart, Belgium.
Gause, GW. 1966. Silvical Characteristics of Bigcone
Douglas-fir (Pseudotsuga macrocarpa (Vasey) Mayr).
USDA Forest Ser- vice PSW Forest & Range
Experiment Station, Research Paper PSW–31,
Berkeley, California.
Gaussen, H. 1955. Parenté réelle et pseudoparenté de
conver- gence. Comptes Rendu Hebdomadaires des
Séances de l’Aca- démie des Sciences, Paris
241(24):1678–1680.
304 Douglas-fir: The Genus Pseudotsuga
61(6):283–296. Goebel, K. 1905. Organography of Plants. Vol. II. Oxford Uni-
Gergácz, J, and C Mátyás. 1993. A VEAB területén versity Press, Oxford.
folytatott fenyötelpitések adatai, áttekintö értékelése Goes, E. 1991. Pseudotsuga (Pseudotsuga menziesii Franco),
(Overview and data of comparative conifer trials in pp. 99–104 in A floresta portuguesa, sua importância e
Transdanubia). Erdészeti Lapok 128. descrição das espécies de maior interesse. Portucel,
Gerhold, HD. 1966. Growth and winter injury of Douglas-fir Lisbon, Portugal.
in a three–year–provenance demonstration. 13th Goldfarb, B, GT Howe, LM Bailey, SH Strauss, and JB
Northeastern Forest Tree Improvement Conference Zaerr. 1991. A liquid cytokinin pulse induces
Proceedings 13:50–52. adventitious shoot formation from Douglas-fir
Gessel, SP. 1980. Regional Forest Nutrition Research cotyledons. Plant Cell Reports (Online), 10:156–160.
Project Biennial Report 1978–1980. Institute of Forest Goldfarb, B, EE Nelson, and EM Hansen. 1989a.
Resources Contribution 39, College of Forest Resources, Trichoderma spp.: Growth rates and antagonism to
University of Washington, Seattle. 44 p. Phellinus weirii in vitro. Mycologia 81(3):375–381.
Gessel, SP, RM Kenady, and WA Atkinson (eds.). 1979. Pro- Goldfarb, B, EE Nelson, and EM Hansen. 1989b.
ceedings of the Forest Fertilization Conference, Trichoderma species from Douglas-fir stumps and roots
September 25-27, Union, Washington. Institute of Forest infested with Phellinus weirii the western Cascade
Resources Contribution Number 40, College of Forest Mountains of Oregon. Mycologia 81:134-138.
Resources, University of Washington, Seattle. Online: Gomes, MA, and F Raposo. 1939. Estudos dendrológicos. I.
http://www. O género Pseudotsuga no Parque da Pena (Sintra). Anais
cfr.washington.edu/research.smc/rfnrp/1FFC_Intro1.pdf do Instituto Superior de Agronomia 10:121–135.
Gessel, SP, EC Steinbrenner, and RE Miller. 1979. Response Gomis, H. 1974. (Establishment of a plantation of
of Northwest forests to elements other than nitrogen. pp. Pseudotsuga menziesii var. viridis in the southern
140–149 in Proceedings of the Forest Fertilization Cordillera of Rio Negro territory, Argentina). Revista
Conference, September 25-27, Union, Washington, ed. SP Forestal Argentina 18(3):73–75. Original not seen, cited
Gessel, RM Kenady, and WA Atkinson. Institute of from Forestry Ab- stracts 36:2651, 1975.
Forest Resources Contribution Number 40, College of
Gorrie, RM. 1965. The Taymount plantation of 1860–1923.
Forest Resources, University of Washington, Seattle.
Online: http://www. Scottish Forestry 19(2):137–140.
cfr.washington.edu/research.smc/rfnrp/1FFC_Intro1.pdf Govedar, Z, D Ballian, T Mikić, and K Pintarić. 2003.
Giacobbe, A. 1942. Econviente la cultura della Douglasia Uspijevan- je različitih provenijencija zelene duglazije
verde (Pseudotsuga Douglasii Carr. var. viridis) in (Pseudotsuga menziesii Mirb. Franco) u okviru IUFRO
Italia? Revista Forestale Italiana 4(4):9–25. programa na oglednoj površini crna ‘Lokva’ kod
Gradiške. Šumarstvo 55(3-4):61–74. [Success of different
Giacobbe, A. 1963. La Douglasia e il pino strobo in Italia. Douglas fir provenanc- es (Pseudotsuga menziesii (Mirb.)
Accademia Italiana di Scienze Forestali, Annali 12:111–144. Franco) in the scope of IUFRO programme on
Giacobbe, A. 1967. La Douglasia nell’ Appennino. Annali Investigation of Douglas-Fir Prov- enance in the test
Ac- cademia Italiana di Scienze Forestali 16:353–380. area “Crna lokva” near the Gradiška.]
Giertych, M. 1967. Analogy of the differences between Gosling, PG. 1988. The effect of moist chilling on the subse-
male and female strobiles in Pinus to the differences quent germination of some temperature conifer seeds
between long- and short-day plants. Canadian over a range of temperatures. Journal of Seed
Journal of Botany 45:1907–1910. Technology 12:90–98.
Giertych, M. 1978. Role of light and temperature in Gosling, PG, Y Samuel, and A Peace. 2003. The effect of
flowering of forest trees. Proceedings of the 3rd mois- ture content and prechill duration on dormancy
FAO/IUFRO World Consultation Forest Tree Breeding, breakage of Douglas-fir seeds (Pseudotsuga menziesii
2:1013–1021. var. menziesii [Mirb.] Franco). Seed Science Research
Giertych, M. 1987. Seed orchards in crisis. Forest Ecology and 13:239–246.
Management 19:1–7. Graff, JE, RK Hermann, and JB Zaerr. 1999. Dry matter and
Giertych M, and Z Królikowski. 1978. Importance of bud nitrogen allocation in western redcedar, western
insolation on female flower induction in pine (Pinus hemlock, and Douglas-fir seedlings grown in low- and
silvestris L.). Arboretum Kórnickie 23:161–169. high-N soils. Annals of Forest Science 56:529–538.
Gifford EM, and AS Foster. 1988. Morphology and Graham, JH. 1979. The role of ethylene in ectomycorrhiza
Evolution of Vascular Plants. WH Freeman and formation and Fusarium infection of Douglas fir roots.
Company, New York, p. 311. Doctoral dissertation, Oregon State University,
Gill, RS. 1981. Factors affecting nitrogen nutrition of Corvallis. Online: http://hdl.handle.net/1957/42928.
western hemlock. PhD dissertation. Oregon State Graham, JH, and RG Linderman.1981. Effect of ethylene
University, Corvallis. 98 p. on root growth, ectomycorrhizal formation and Fusar-
Gill, R, and DP Lavender. 1983a. Urea fertilization: ium infection of Douglas-fir. Canadian Journal of
effects on primary root mortality and mycorrhizal Botany 59(2):149–155.
development of young growth western hemlock. Forest Graham, JH, and RG Linderman. 1983. Pathogenic seed
Science 29:751–760. borne Fusarium oxysporum from Douglas-fir. Plant Dis-
Gill, RS, and DP Lavender. 1983b. Urea fertilization and ease 67:323–325.
foliar nutrient composition of western hemlock [Tsuga Graham JS, SD Hobbs, and JB Zaerr. 1994. The effect of
heterophylla (Raf.) Sarg.]. Forest Ecology and flurprimidol on bud flush, shoot growth, and on en-
Management 6:333–341. dogenous gibberellins and abscisic acid of Douglas-fir
Gilmore, AR. 1965. The Apparent Source of a Growth seedlings. Journal of Plant Growth Regulation 13(3):131–
Stimulus in Loblolly Pine Seedlings. Forest Note #112, 136.
Department of Forestry, Illinois Agriculture Experiment Graham, K, and Prebble. 1941. Investigations of the Role of
Station, Uni- versity of Illinois. 4 p. Insects in Relation to Seed Production in Douglas-fir.
Glover, MM, JR Sutherland, CL Leadem, and G Shrimpton. Annual For- est Insect Investigations and Detailed
1987. Efficiency and Phytocity of Fungicides for Control Studies in Coastal British Columbia. Canadian
of Botrytis Grey Mold on Container-Grown Conifer Department of Agriculture, Forest Biology Division,
Seedlings. FRDA Report 012 Economic and Regional Victoria, BC. 21 pp.
Development Agreement. Canadian Forestry Service, Graham, K, and ML Prebble. 1942. Investigations of the Role
British Columbia Ministry of Forests. of Insects in Relation to Seed Production in Douglas-fir.
An-
References 305

nual Forest Insect Investigations and Detailed Studies Southern Arkansas and Northern Louisiana. USDA
in Coastal British Columbia. Canadian Department of Forest Service Research Note SO–92, Southern Forest
Agriculture, Forest Biology Division, Victoria, BC. 9 Experi- ment Station, New Orleans, LA.
p. Grime JP, JC Crick, and JE Rincon. 1986. The ecological
Gratkowski, H.J. 1956. Windthrow around staggered settings signifi- cance of plasticity. Symposium of the Society for
in old-growth Douglas-fir. Forest Science 2:60–74. Experimental Biology 40, 5–29.
Gravatt, AR, DH Lalham, LWR Jackson, GY Young, and Grob, JA. 1990. Techniques to Study the Cell Cycle in the Shoot
WC Davis. 1940. Multiple seedlings of pines and Apex of Conifers. M. Science thesis. University of Victoria,
Douglas-fir. Journal of Forestry 38:818. Victoria, BC.
Gray, AN, and JF Franklin. 1997. Effects of multiple fires on Grob, JA, and JN Owens. 1994. Techniques to study the cell
the structure of southwestern Washington forests. cycle in conifer shoot apical meristems. Canadian
Northwest Science 71(3):174–184. Journal of Forest Research 24(3):472.
Greacon, EL, and R Sands. 1980. Compaction of forest soils: Groome, JG. 1978. Position Paper—JG Groome and
A review. Australian Journal of Soil Research 18:163– Associates, pp. 355–357 in A Review of Douglas-fir in
189. New Zealand. FRI Symposium No. 15 Forest Service,
Greenwood, MS. 1987. The role of dormancy in the develop- Forest Research Institute.
ment of male and female strobilis of lobloll pine. Forest Groos. 1968. Die Douglasie–Neubürgerin der hessischen
Science 23(3):373–375. Wälder. Wald Forum ‘68 Sonderausgabe des
Greenwood, MS. 1992. Theoretical aspects of juvenility and Staatsanzeiger für das Land Hessen:49–51.
maturation. pp. 19–25 in Proceedings of the IUFRO Sym- Grossnickle, S, and RS Folk. 1993. Stock quality assessment
posium, Mass production technology for genetically forecasting survival or performance on a reforestation
improved fast growing forest tree species, Bordeaux, site. Tree Planters Notes Summer 114–121.
September 1992. AFOCEL, Nancy. Groth, P. 1927. Die Wurzelbildung der Douglasie und ihr
Greenwood, MS. 1995. Juvenility and maturation in conifers: Ein- fluss auf die Sturm und Schneefestigkeit dieser
current concepts. Tree Physiology 15:433–438. Holzart. Allgemeine Forst-und Jagdzeitung 103:106–205,
Greenwood, MS, and KW Hutchison. 1993. Maturation as 217–231,
a developmental process, pp. 14–33 in Clonal forestry. I. 268–285.
Genetics and Biotechnology, ed. MR Ahuja and WJ Groth, O. 1928. Beitrag zur Frage der Sturmfestigkeit der
Libby. Springer-Verlag, New York. grünen Douglasie. Allgemeine Forst-und Jagdzeitung
Greenwood, MS, and KW Hutchison. 1996. Genetic aftereffects 104:324–226.
of increased temperature in Larix, pp. 56–62 in Grundner, F. 1921. Die Anbauversuche mit fremdländisch-
Proceedings of the 1995 Meeting of the Northern Global en Holzarten in den braunschweigischen Staatsforsten.
Change Program, ed. J Hom, R Birdsey and K O’Brian. Mitteilungen der Deutschen Dendrologischen Gesellschaft
USDA Forest Service Report NE-214 . 31:19–68.
Greenwood, MS, GW Adams, and MG Gillespie. 1991. Guak S, Olsyzk, DM, Fuchigami LH, DT Tingey. 1998.
Stimu- lation of flowering by black spruce and white Effects of elevated CO2 and temperature on cold
spruce: a comparative study of the effects of gibberellins hardiness and spring bud burst and growth in
A4/7, Cultural treatments, and environment. Canadian Douglas-fir (Pseudotsuga
Journal of Forestry Research, 21:395–400. menziesii). Tree Physiology 18(10):671–679.
Grier, CC, and RH Waring. 1974. Conifer foliage mass Gugger, PF, and S Sugita. 2010. Glacial populations and
related to sapwood area. Forest Science 20:205–206. postglacial migration of Douglas-fir based on fossil
Griffin, AR, and AC Matheson. 1978. Early height growth of pol- len and macrofossil evidence. Quaternary Science
some provenances from the IUFRO Douglas-fir collec- Reviews 29: 2052–2070.
tion in nine Australian field trials, pp. 217–228 in Vol. 1: Gugger PF, Sugita S, Cavender-Bares J. 2010.
Background papers and Douglas-fir provenances. Pro- Phylogeography of Douglas-fir based on mitochondrial
ceedings IUFRO joint meeting of Working Parties S2.02– and chloroplast DNA sequences: testing hypotheses
05 Douglas-fir provenances, S2.02–06 Lodgepole pine from the fossil record. Molecular Ecology 19:1877–
provenances, S2.02–012 Sitka spruce provenances, S2–14 1897.
Abies provenances. Vancouver, BC. 1978. BC Ministry of Gugger, PF, A González-Rodríguez, H Rodríguez-Correa,
Forests, Information Service Branch, Victoria, BC. S Sugita, and J Cavender-Bares. 2011. Southward
Griffin, AR. 1978. Geographic variation in Douglas-fir from Pleisto- cene migration of Douglas-fir into Mexico:
the coastal ranges of California. II. Predictive value of a phylogeog- raphy, ecological niche modeling, and
regression model for seedling growth variation. Silvae conservation of ‘rear edge’ populations. New
Genetica 27 (3/4):96–101. Phytologist 189: 1185–1199.
Griffin, GJ. 1919. Bordered pits in Douglas-fir. A study of Günzl, L. 1981. Ergebnisse aus den österreichischen
the position of the torus in mountain and lowland Douglasien– Provenienzversuchen. Forstliche
specimens in relation to creosote penetration. Journal Bundesversuchsanstalt Wien, Informationsdienst 203.
of Forestry 17(11):813–822. Folge.
Griffin, JR. 1964. A new Douglas-fir locality in Southern Günzl, L. 1986. Anbauerfahrungen aus den
California. Forest Science 10(3):317–319. österreichischen Douglasien–Provenienzversuchen der
Griffin, RA, and KK Ching. 1977. Geographic variation in letzten 20 Jahre. Reprint from Allgemeine
Douglas-fir from the coastal ranges of California. I. Forstzeitung (Wien) Folge 2, Feb. 1986.
Seed, seedling growth and hardiness characteristics. Günzl, L. 1987. Douglasien–Anbauversuche in Österreich.
Silvae Genetica 26(5/6):149–157. Der Förderungsdienst 35(11):316–326.
Griffith, BG. 1968. Phenology, Growth, and Flower and Cone Gur, A. 1985. Rosaceae-deciduous fruit trees, p. 355–390. In
Pro- duction of 154 Douglas-fir Trees on the University A Halevy (ed.). Handbook of Flowering, Vol. I. CRC,
Research Forest as Influenced by Climate and Fertilizer, Bota Raton, FL.
1957–1967. University of British Columbia Faculty of Guttenberg, H. von. 1941. Der primäre Bau der
Forestry Bul- letin No. 6, Vancouver, BC. 70 p. Gymnosper- menwurzel. Handb. der
Grigsby, HC. 1969. Exotic Trees Unsatisfactory for Forestry in Pflanzenanatomie. VIII. Berlin.
Haase, DL, and R. Rose. 1993. Soil moisture stress induces
transplant shock in stored and unstored 210 Douglas-
fir seedlings of varying root volumes. Forest
Science 39(2):275–293.
306 Douglas-fir: The Genus Pseudotsuga
Hacke-Oudemans, JJ, and TC Oudemans. 1955. Herkomst California.
van de douglas in Nederlanden. Nederlands bosbouwti- Hansen, EM, LF Roth, PB Hamm, and AJ Julis. 1980. Sur-
jdschrift 27(4):81–88.
Hackett, WP. 1985. Juvenility, maturation, and rejuvenation
in woody plants. Horticultural Reviews 7:109–155.
Hackett, WP, and JR Murray. 1992. Propagation of
ornamental
plants. Acta Horticulturae 314:195–203.
Hackett, WP, JR Murray, and A Smith. 1992 Control of
matura- tion in woody species, pp. 45-50 and 83–90 in
Proceedings of the IUFRO Symposium, Mass
Production Technology for Genetically Improved Fast
Growing Forest Tree Species, Bor- deaux, 14-18
September 1992. AFOCEL, Nangis.
Hacskaylo, E. 1972. Mycorrhiza: The ultimate in reciprocal
parasitism. BioScience 22(10):577–583.
Hacskaylo, E. 1973a. Carbohydrate physiology of ectomy-
corrhizae, pp. 207–230 in GC Marks and TT Kozlowski,
Ectomycorrhizae: Their Ecology and Physiology.
Academic Press, New York. 444 p.
Hacskaylo, E. 1973b.Dependence of mycorrhizal fungi on
hosts. Bulletin of the Torrey Botanical Club 100(4):217–
223.
Haddock, PG, J Walters, and A Kozak. 1967. Growth of Coastal
and interior provenances of Douglas-fir (Pseudotsuga
menziesii (Mirb.) Franco) at Vancouver and Haney in
British Columbia. Research Paper, Faculty of Forestry
University of British Columbia No. 79, Vancouver, BC.
Hagar, DC. 1960. The interrelationships of logging birds
and timber regeneration in the Douglas fir region of
north- western California. Ecology 4(1):116–125.
Hagem, O. 1931. Forsøk med vestamerikanske traeslag (Ex-
periment with west American tree species). Vestlandets
forstlige forsøksstasjon 4(12):1–217.
Hagenstein, WD. 1986. Douglas-fir forestry from Alien to
Mendel, pp. 32–38 in Douglas-fir: Stand management for
the future, ed. CD Oliver, DP Hanley, and JA Johnson.
College of Forest Resources, University of
Washington, Seattle.
Hagman, M. 1973. Development in the nursery of the Finnish
part of the IUFRO experiment with Pseudotsuga
menziesii (Mirb.) Franco, pp. 216–226 in Proceedings
Meeting of IU- FRO Working Party on Douglas-fir
provenances S2.02–05, Sept.3–5, Göttingen, Germany.
HH Hattemer ed.
Haig, IT, Davis, KP, and RW Wredman. 1946. Natural
Regen- eration in the Western White Pine Type. USDA
Technology Bulletin No. 76. Washington, DC. 98 p.
Haig, IT. 1936. Factors controlling initial establishment of
western white pine and associated species. Yale
University School of Forestry Bulletin 41.
Haig, IT, KP Davis, and RH Weidman. 1941. National Re-
generation in the Western White Pine Type. USDA Tech
Bulletin 767.
Hall, RE. 1955. Insect Damage to the 1954 Crop of Douglas-fir
and Sugar Pine Cones and Seeds in Northern California.
California Miscellaneous Paper #18. California Forest and
Range Experiment Station, p. 2.
Halliday, WED, and AWA Brown. 1943. The distribution of
some important forest trees in Canada. Ecology
24(3):353– 373.
Hamlin, J. 1990. Genetic structure of Douglas-fir populations
from Southwestern Oregon. Paper 2.128, pp. 1–16 in
Pro- ceedings Joint Meeting WFGA and IUFRO WP’s
S2.02, 05, 06, 12, 14. Olympia, WA, Weyerhaeuser Co.,
Centralia, WA.
Hamrick, JL, JB Mitton, and YB Linhart. 1979. Levels of
genetic variation in trees: influence of life history
characteristics, pp. 35–41 in Proceedings of Symposium on
Isozymes of North American Forest Trees and Forest
Insects, coord. MT Conkle. USDA, Forest Service
General Technical Report PSW–48, Pacific Southwest
Forest and Range Experiment Station, Berkeley,
vival, spread, and pathogenicity of Phytophthora Doug- lasien (Pseudotsuga menziesii var. caesia (Schwer.)
spp. on Douglas-fir seedlings planted on forest Franco) im Sächsischen Erzgebirge, pp. 82–88 In
sites. Phytopathol- ogy 70(5):422–425. Beiträge zur Gehölzkunde 1997. Verlag Gartenbild
Hansen, EM, JK Stone, BR Capitano, P Rosso, W Heinz Hausmann, Rinteln, Germany.
Sutton, L Winton, A Kanaskie, and MG Hartig, R. 1892. Über die bisherigen Ergebnisse der
McWilliams. 2000. Incidence and impact of Swiss Anbauver- suche mit ausländischen Holzarten in den
needle cast in forest plantations of Douglas-fir in bayrischen Sta-
coastal Oregon. Plant Diseases 84:773–778.
Hanson, HS. 1952. Megastigmus seed flies. Quarterly
Journal of Forestry 46(4):261–264.
Haralamb, A. 1971. Cultura speciilor forestiere. Ed. III–a.
Editura Agro–Silvica, Bucharest.
Harkai, L. 1971. A magyarorszagi duglásfenyö–
állományok termöhelyi és fatermési vizsgálata. I.
rész. A termöhelyi igény vizsgálata (Site and
yield examination of Douglas- fir stands in
Hungary. Part I. The examination of site
requirements). Erdészeti Kutatások 67(1):149–
168.
Harkai, L. 1975. Yield and improvement experiments
of Douglas-fir in Hungary. Erdészeti Kutatások
70(2):37–46.
Harkai, L. 1981. A zalaerdödi fenyö hálózati
kisérletek ér- tékelése (Evaluation of the conifer
comparative trials at Zalaerdöd). Erdészeti
Kutatások 74:89–96.
Harkai, L. 1983. A Zalaerdödi duglaszfenvö–
Szarmazasi kiserlet fatermestani ertekelese (The
evaluation of the yield of the Douglas-fir
provenance trial at Zalaerdöd). Erdeszeti
Kutatasok 75(1):19–27.
Harkai, L. 1987. A duglászfenyö hálózati kisérlet
értékelése (The evaluation of a spacing
experiment with Douglas- fir). Erdészeti
Kutatások 79:33–38.
Harlow, WM, and ES Harrar. 1969. Textbook of Dendrology, 5th
ed. McGraw–Hill Book Co., New York.
Harrer, F. 1925. Die Douglasfichte in Bayern nach dem
Stande ihres Anbaus am 1. Januar 1923.
Mitteilungen der Staats- forstverwaltung Bayerns
16:102–209.
Harrington, CA, and DL Reukema. 1983. Initial
shock and long-term stand development
following thinning in a Douglas-fir plantation.
Forest Science 29(1):33–46.
Harrington, JF. 1972. Seed storage and longevity. In
Seed Bi- ology, Vol III. Insects, and Seed Collection,
Storage, Testing, and Certification, ed. TT
Kozlowski. Academic Press, New York, pp. 145–
245
Harrington, M. 1991. Fire management in interior
Douglas- fir forests, pp. 209–214 in Proceedings of
a Symposium on Interior Douglas-fir, the species
and its management, ed. D Baumgartner and J
Lotan. Cooperative Extension, Wash- ington State
University, Pullman.
Harrington, TB. 1989. Stand Development and
Individual Tree Morphology and Physiology of
Young Douglas- fir (Pseudotsuga menziesii) in
Association with Tanoak (Lithocarpus
densiflorus). Doctoral dissertation, Oregon State
University, Corvallis.
Harrington, TB, RJ Pabst, and JC Tappeiner II. 1994.
Seasonal physiology of Douglas-fir saplings:
response to micro- climate in stands of tanoak or
Pacific madrone. Forest Science 40:59–82.
Harris, AS. 1971. Experience with Douglas-fir in Southeast
Alaska. Northwest Science 45(2):87–93.
Hartig, M. 1980. Gegenwärtiger Stand des Anbaus
fremdlän- discher Baumarten im Mittelgebirge
und Hügelland der DDR, pp. 69–80 in Kolloquium
Tharandt–Brno, June 1980. Teil 3. Technische
Universität Dresden, Sektion Forst- wirtschaft
Tharandt.
Hartig, M. 1997. Wuchsverhalten und Vitalität Grauer
References 307
Hedlin. AF. 1959b. Description and habits of a new species of
atswaldungen. Forstlich–Naturwissenschaftliche Zeitschrift
1(11):401–432.
Harvey, AE, MF Jorgensen, MJ Lansen, and RT Graham.
1987. Decay Organic Material and Soil Quality in the
Inland Northwest. A Management Opportunity, USDA
Forest Service Int. General Technical Report 225, 15 p.
Harvey, GM, and LR Carpenter. 1945. Fungi on stored
Doug- las-fir cones–a problem? Tree Planters’ Notes
26(4).
Haskell, AR, and E Voss. 1954. The pharmacology of tetramine
(Tetraethylene disulfate tramine). Journal of the American
Pharmaceutical Association Vol. XLVI No. 4
Hattemer, HH, and A König. 1975. Geographic variation
of early growth and frost resistance in Douglas-fir.
Silvae Genetica 24(4):97–106.
Hausrath. 1921. Erfahrungen mit dem Anbau fremder Hol-
zarten in Baden. Mitteilungen der Deutschen
Dendrologischen Gesellschaft 31:242–244.
Havraneck WM, Tranquillini W. 1995. Physiological
processes during winter dormancy and their ecological
significance, pp. 95–124 in Ecophysiology of Coniferous
Forest, ed. WK Smith and TM Hinckley, Academic
Press, New York.
Hawkins, CDB, and WD Binder. 1990. State of the art
seedling stock quality tests based on seedling
physiology, pp. 91–121 in The Target Seedling
Symposium, ed. R Rose, SJ Campbell and TD Landis.
USDA Forest Service, General Technical Report RMA
200, Fort Collins, CO.
Hawkins, CDB, and GR Lister. 1985. In vivo chlorophyll
fluorescence as a possible indicator of the dormancy
stage in Douglas-fir seedlings. Canadian Journal of
Forest Research 15(4):607–612.
Hawkins, BJ, HJ Guest, and D Rolotelo. 2003. Freezing
toler- ance of conifer seeds and germinants. Tree
Physiology 23:1237–1246.
Hawkins, BJ, SBR Kiiskila, and G Henry. 1999. Biomass
and nutrient allocation in Douglas-fir and amabilis fir
seed- lings: influence of growth rate and temperature.
Tree Physiology 19:59–63.
Hawksworth, FG, and D Wiens. 1972. Biology and
classifica- tion of dwarf-mistletoes (Arceuthobium). USDA
Agriculture Handbook 401. Washington, DC. 234 p.
Hayashi, Y. 1952. The Natural Distribution of Important Trees
Indigenous to Japan. Conifers Report 2. Tokyo Forest Ex-
periment Station Meguro, Bulletin 55.
Hayata, B. 1905. On the distribution of the formosan
conifers.
Tokyo Botanical Magazine 9:43–60.
Hayata, B. 1915. Icones Planatarum Formosanarum. Vol. 5. Bureau
Prod. Ind. Gov. Formosa, Taihoka.
Hedderwick, GW. 1968. Prolonged drying of stratified
Douglas-fir seed affects laboratory germination. Research
leaflet #19. 2A. Forest Research Institute, New Zealand
Forest Service, Rotorua.
Hedgcock, GG, GF Gravatt, and RP Marshall. 1925.
Polyporus schweinitzii Fr. on Douglas-fir in the eastern
United States. Phytopathology 15:568–569.
Hedlin, AF. 1958a. Studies on Cone and Seed Insects in
British Columbia. Interim Report. 1958. Forest Biology
Lab Research Branch, Forest Biology Division Canada
Department of Agriculture. 13 p.
Hedlin, AF. 1958b. Studies on cone and Seed Insects in
British Columbia. Forest Biology Lab Research Branch,
Forest Biology Division Canada Department of
Agriculture. 20 p.
Hedlin, AF. 1959a. The effect of moisture and temperature
on the emergence of the larvae of the Douglas-fir cone
midge, Contarinia oregonensis Foote from cone scales.
For- est Biology Lab Research Branch, Forest Biology
Division Canada Department of Agriculture. Bi-monthly
Progress Report 15:3–4.
N. America. Proceedings of the Entomological Society of
Phytophaga (Diptera: Cecidomyiidae) from western red Wash- ington 12(4):191–197.
cedar cones. The Canadian Entomologist 91(11):719–723. Heikinheimo, O. 1956. Tuloksia ulkomaisten puulajien viljely-
Hedlin, AF. 1960. On the life history of the Douglas-fir
cone moth, Barbara colfaxiana (Kft.) Lepidoptera:
Olethreuti- dae) and one of its parasites, Glypta
evetriae Cush. (Hy- menoptera: lchneumonidae.
The Canadian Entomologist 92(11):826–834.
Hedlin, AF. 1961a. The life history and habits of a
midge, Contarinia oregonensis Foote (Diptera:
Ceidomyiidae) in Douglas-fir cones. The Canadian
Entomologist 93:952–967.
Hedlin, AF. 1961b. Some aspects of the cone and seed
in- sect problem in the Pacific Northwest. Forestry
Chronicle 32(4):6–9.
Hedlin, AF. 1962. Two systemic insecticides
Phosphamidon and Systox used against the
Douglas-fir cone midge, Contarinia oregonensis
Foote. Canadian Department of Agriculture, Forest
Biology Division. Bi Monthly Progress Report
18(1):3–4.
Hedlin, AF. 1964a. A six-year plot study on Douglas-fir
cone insect population fluctuations. Forest Science
10(1):124–128.
Hedlin, AF. 1964b. Five systemic insecticides used
against Douglas-fir cone insects. Bi-monthly
Progress Report. Ca- nadian Department of
Forestry 20(2):4.
Hedlin, AF. 1966. Prevention of insect-caused seed loss
in Douglas-fir with systemic insecticides. Forestry
Chronicle, Quebec 42(1):76–82.
Hedlin, AF. 1974. Cone and Seed Insects of British
Columbia. Environment Canada, Info Report BC-
X-90, Canadian Forestry Service Pacific Forest
Research Center, Victoria, BC. 63 p.
Hedlin, AF, and NE Johnson. 1963. Life history and
habits of a midge, Contarinia washingtonensis
Johnson (Diptera: Cecidomyiidae) in Douglas-fir
cones. The Canadian En- tomologist 95:1168–1175.
Hedlin, AF, and NE Johnson. 1968. A new species of
Camp- tomyia (Diptera: Cecidomyiidae) from
Douglas-fir cones. The Canadian Entomologist
100(5):532–535.
Hedlin, AF, and DS Ruth. 1968. Sex attraction in the
Douglas-fir cone moth Barbara colfaxiana (Kft.)
Canadian Department of Agriculture, Forest Biology
Division Bi-monthly Progress Report 24(1):7–8.
Hedlin, AF, and DS Ruth. 1978. Examination of
Douglas-fir clones for differences in susceptibility
to damage by cone and seed insects. Journal of the
Entomological Society of British Columbia 75:33–
34.
Hedlin, AF, GE Miller, and DS Ruth. 1982. Induction of
pro- longed diapause in Barbara colfaxiana
(Lepidoptera: Ole- threutidae): Correlations with
cone crops and weather. The Canadian
Entomologist 114(6):465–471.
Hedlin, AF, J Weatherston, DS Ruth, and GE Miller.
1983. Chemical lure for male Douglas-fir cone
moth, Barbara colfaxiana (Lepidoptera:
Olethreutidae). Environmental Entomology
12(6):1751–1753.
Hedlin, AF, HO Yates III, DC Tovar, BH Ebel, TW
Koerber, and EP Merkel. 1980. Cone and Seed
Insects of North Ameri- can Conifers. Environment
Canada, Canadian Forestry Service, Ottawa, ON;
USDA Forest Service, Washington, DC; Secretaria
de Agricultura y Recursos Hidraulicos, Chapingo,
Mexico. 122 p.
Heiberg, HHH. 1975. North American trees in Norway.
The Forestry Chronicle 51(5):183–184.
Heiberg, HHH. 1978. Douglasgranen i Norge (with
English summary: The Douglas fir in Norway).
Tidsskrift for Sk- ogbruk 86(3):147–160.
Heidemann, O. 1910. New Species of Leptoglossus from
308 Douglas-fir: The Genus Pseudotsuga
stä Suomessa. (with German summary: Ergebnisse Hermann, RK. 1977. Observations on frost damage to Douglas- fir
von einigen Anbauversuchen mit fremdländischen Christmas trees in Oregon. American Christmas Tree
Holzarten in Finnland). Communicationes Instituti
Forestalis Fenniae 45(3):1–129.
Heikinheimo, O. 1957. Ulkomaiset puulajit Suomen metsäta-
loudessa. (with English summary: Exotic tree species in
Finnish forestry). Paperi ja Puu 39(4a):211–216.
Heilman, P. 1982. Response Potentials to Elements Other
Than Nitrogen Stand Management Workshop: Fertility.
Forestry Extension Service, Department of Forest
Science, Oregon State University, Corvallis, OR.
Heilman, PE, and G Ekuan. 1973. Response of Douglas-fir
and western hemlock seedlings to lime. Forest
Science 19(3):220–224.
Heilman, PE, HW Anderson, and DM Baumgartner, eds.
1979. Forest Soils of the Douglas-fir Region.
Washington State University Cooperative Extension
Service, Washington State University, Pullman. 298 p.
Heineman, JL and DP Lavender. Growth of interior spruce
seedlings on forest floor materials. Unpublished ms.
19 p.
Heiner, TD, and DP Lavender. 1972. Early growth and
drought avoidance in Douglas-fir seedlings. Research
Paper 14, Forest Research Laboratory, School of
Forestry, Oregon State University, Corvallis. 7 p.
Hejtmanek, J. 1952. Klimatické moznosti rozsíreni duglasky
zelené u nás (Climatic possibilities of extending the cul-
tivation of Pseudotsuga taxifolia var. viridis in
Czechoslo- vakia). Lesnicka práce 31(3):100–107.
Helms, JA. 1964. Apparent photosynthesis of Douglas-fir in
relation to silvicultural treatment. Forest Science
10(4):432– 442.
Helms, JA. 1965. Diurnal and seasonal patterns of net
assimila- tion in Douglas-fir (Pseudotsuga menziesii
(Mirb.) Franco, as influenced by environment. Ecology
46(5):698–707.
Hellmers, H. 1963. Effects of soil and air temperature on
growth of redwood seedlings. Botannical Gazette
125:172–177.
Henderson, J, and W Brubaker. 1986. Response of Douglas-
fir to long-term variation in precipitation and
temperature in western Washington, pp. 162–167 in
Douglas fir: Stand Management for the Future, ed. CD
Oliver, DP Hanley, and JA Johnson. Institute of Forest
Resources Contribution No. 55, University of
Washington, Seattle.
Heninger, RL, and DP White. 1974. Tree seedling growth
at different soil tempentures. Forest Science 20:363–
367.
Henkel, W. 1960. Untersuchungen uber die Sturmkatastro-
phe im staatlichen Forst wirtschaftsbetrieb Sonneberg/
Thüringen in 1958. Archiv für Forstwesen 9(1):28–46.
Henne, A. 1970. Douglasien–Anbau in Hessen.
Allgemeine
Forstzeitschrift 25(39):808–810.
Henrikson, HA. 1956. Sitka–Fichte und Douglasie in der
daenischen Forstwirtschaft. Allgemeine
Forstzeitschrift 11(45/46):581–583.
Henry, A, and MG Flood. 1920. The Douglas firs: A botani-
cal and silvicultural description of the varies species of
Pseudotsuga, pp. 67–92 in Royal Irish Academy
Proceedings 35, Section B, No. 5.
Hepting, George H. 1971. Diseases of forest and shade trees of
the United States. US Department of Agriculture,
Agriculture Handbook 386. Washington, DC. 658 p.
Hermann, RK. 1967. Seasonal variation in sensitivity of
Douglas-fir seedlings to exposure of roots. Forest Sci-
ence 13(2):140–149.
Hermann, RK. 1974. Frost damage in the nursery and its
effect of field performance of Douglas-fir seedlings.
Western Forest Nursery Council Meeting Portland,
Oregon, Au- gust 5–7, 1974. Proceedings:117a–117d.
Journal 21(1):13–19. Göttingen, Germany. ed. HH Hattemer.
Hermann, RK. 1977. Growth and production of tree Hessische Forstliche Versuchsanstalt. 1993. Douglasien–
roots, in The below-ground ecosystem: a synthesis of Einzelstammabsaatenversuch (Aussaat 1970). Jahresb-
plant-associated processes. ed. JK Marshall, ericht 1993:8.
Colorado State University, Range Science
Department, Science Series 26. Fort Col- lins, CO.
351 p.
Hermann, RK. 1978. Reproductive systems p. 27–38 in
Regener- ating Oregon’s Forests, ed. BD Cleary,
RD Greaves, and RK Hermann. OSU Extension
Service, Corvallis, OR. 287 p.
Hermann, RK. 1982.The genus Pseudotsuga: Historical
Records and Nomenclature. Special publication 2a.
Forestry Re- search Laboratory, School of Forestry,
Oregon State Uni- versity, Corvallis.
Hermann, RK. 1985. The Genus Pseudotsuga: Ancestral
His- tory and Past Distribution. Special
Publications 2b, Forest Research Laboratory,
Oregon State University, Corval- lis. 32 p.
Hermann, RK. 1987. North American tree species in Europe.
Journal of Forestry 85 (12):27–32.
Hermann, RK. 1990. Cold injury; frost damage; frost
heaving; winter desiccation, pp. 68–72 in Growing
healthy seedlings, ed. PB Hamm, SJ Campbell, and
EM Hansen. Published by Forest Pest Management,
USDA Forest Service, Pa- cific Northwest Region,
and Forest Research Laboratory, College of
Forestry, Oregon State University, Corvallis.
Hermann, RK. 2005. Wurzelstudien an Douglasie, Zum
An- bau und Wachstum der Douglasie, PH Dong
(Hrsg.). Mit- teilungen aus der Forschungsanstalt
für Waldökologie und Forstwirtschaft Rheinland-
Pfalz 55:135–164. [RK Hermann, 2005. Studies of
the root system of Douglas fir, pp. 135–164 in
About cultivation and growth of Douglas-fir, ed. PH
Hong, Rheinland-Pfalz Ministry of Environment
and Forests 55.]
Hermann, RK, and WW Chilcote. 1965. Effect of
seedbeds on germination and survival of Douglas-
fir. Oregon State Uni- versity, School of Forestry,
Forest Research Laboratory Research Paper 4.
Corvallis. 28 p.
Hermann, RK, and DP Lavender. 1968. Early growth of
Doug- las-fir from various altitudes and aspects in
southern Oregon. Silvae Genetica 17(4):143–151.
Hermann, RK, and DP Lavender. 1990. Pseudotsuga
men- ziesii (Mirb.) Franco: Douglas-fir, pp. 527–
540 in Silvics of North America: Vol. I, Conifers,
coord. RM Burns and BH Honkala. Agricultural
Handbook 654, USDA Forest Service Washington,
DC.
Hermann, RK, and JB Zaerr. 1973. How to recognize frost
dam- age in forest trees and what to do about it.
Oregon State University, Extension Service. Extension
Circular 820. 4 p.
Hermann, RK, DP Lavender, and JB Zaerr. 1972.
Lifting and Storing Western Conifer Seedlings.
Research Paper 17, Or- egon Forest Research
Laboratory, School of Forestry, Oregon State
University, Corvallis.
Hermann, RK, DP Lavender, and J Zaerr. 1974. Role of
physi- ology in young-growth management:
Nutrition, fer- tilization, irrigation, growth
regulators, pp. 77–104 in Managing Young Forests in
the Douglas-fir Region: Proceed- ings of a
Symposium held June 14-16, 1972. Vol. 4, comp.
and ed. AB Berg, OSU School of Forestry,
Corvallis, OR.
Hernandez, GT, GV Alonso, GP Aribas, and JL Jenkins.
1993. Screening Douglas-fir for Rapid Early Growth
in Common– Garden Tests in Spain. USDA, Forest
Service, General Technical Report PSW–GTR–146.
Herrmann, S. 1973. Preliminary results from Douglas-fir
provenance tests in the Emsland. pp. 37–50 in
Proceedings Meeting of IUFRO Working Party on
Douglas-fir provenances S2.02–05, Sept.3–5,
References 309
Washington, DC. 63 p.
Heusser, CJ. 1960. Late–Pleistocene environments of North Hofmann, JV. 1925. Laboratory tests on effect of heat on seeds of
Pacific North America. American Geographic Society noble and silver fir, western white pine, and Douglas fir.
Special Publication No. 35. Journal of Agricultural Research 31:197–199.
Heusser, CJ. 1968. Discussions: Late Pleistocene coniferous Hofmann, JV, and CP Willis. 1915. A study of Douglas-fir seed.
forests of the Northern Rocky Mountains. p. 22–23 in
Proceedings of the 1968 Symposium, Center for Natural
Re- sources, Missoula, MT, Sept. 19–20, 1968, ed. RD
Taber.
Heyder, JC. 1986. Waldbau im Wandel. JD Sauerlaenders
Verlag,
Frankfurt/M.
Hickel, R. 1922–23. Le sapin de Douglas (Pseudotsuga
douglasii).
Bulletin de la Société. Dendrologique de France 44:51–
79;
45:95–103; 46:5–24.
Hinckley, TM, AL Friend, and AK Mitchell. 1992. Response
at the foliar, tree, and stand levels to nitrogen
fertilization, pp. 82–89, in Forest Fertilization: Sustaining
and Improving Nutrition and Growth of Western Forests,
ed. HN Chapell, GF Weetman, and RE Miller.
University of Washington, Contr. No. 73.
Ho, RH, and JN Owens. 1974. Microstrobili of Douglas-fir.
Canadian Journal of Forest Research 4:561.
Ho, RH, and O Sziklai. 1972. Germination of Douglas-fir
pol- len. Silvae Genetica 21:48–50.
Hobbs, S. 1985. Undercutting affects seedling morphology
but not field performance, pp. 3–4 in Fir Report, OSU
Extension Service, Forestry Intensified Research,
7(4). 16 p. Medford, Oregon.
Hobbs, SD, SD Tesch, PW Houston, RE Stewart, JC Tap-
peiner, and GE Wells, eds. 1992. Reforestation
Practices in Southwestern Oregon and Northern
California. College of Forestry, Forest Research
Laboratory, Oregon State University, Corvallis.
Hocking, D, and RD Nyland. 1971. Cold Storage of Coniferous
Seedlings: a Review. State University College of Forestry
at Syracuse, Applied Forestry Research Institute, Research
Report 6.
Hodges, JD. 1967. Patterns of photosynthesis under
natural
environmental conditions. Ecology 48(2):234–242.
Hoefnagels, MH, and RG Linderman. 1999. Biological sup-
pression of seedborne Fusarium during cold
stratification of Douglas-fir seeds. Plant Disease
83:845–852.
Hofman, J. 1962. Untersuchungen der Douglasienpflanzen
in der CSSR und ihre Resultate. Archiv für Forstwesen
11(12):1208–1318.
Hofman, J. 1964. Pestováni douglasky (with German summary:
Waldbau der Douglasie [in der Tschechoslowakei]). Státhi
zemedelské nakladatelstvi, Prague.
Hofman, J, and B Heger. 1958. Prehled douglaskovch
vysadeb v Ceskoslovensku (List of Douglas-fir stands in
Czecho- slovakia). Zprávy VULH 4(1):91–128.
Hofman, J, M Vackova, and B Heger. 1964. Zpráva o
prvních proveniencuich pokusech s douglaskou
tisolistou v CSSR (with German summary: Nachricht
über die ersten Pro- venienzversuche mit der
Douglasie in der Tschecho- slowakei). Acta Musei
Silesiae, Ser.C, III:43–50.
Hofmann, JV. 1917. Natural Reproduction from Seed Stored
in the Forest Floor. Journal of Agricultural Research 11
(October):1–26.
Hofmann, JV. 1920. The establishment of a Douglas-fir
forest.
Ecology 1(1) 49–53.
Hofmann, JV. 1921. Adaptation in Douglas fir. Ecology
2(2):127–
131.
Hofmann, JV. 1924. The Natural Regeneration of Douglas-fir
in the Pacific Northwest. USDA Bulletin 1200,
sper- motrophus Wachtl (Hymenoptera: Chalcidoidea)
Society of American Foresters Proceedings and its principal parasite, with descriptions of the
developmental stages. Transactions of the Royal
10(2):141–164. Holland, H. 1912. Die Entwicklung Entomological Society of
und der Stand der An-
bauversuche mit fremdländischen Holzarten in den Sta-
atswaldungen Württembergs. Mitteilungen der
Deutschen Dendrologischen Gesellschaft 21:20–54.
(also published in Naturwissenschaftliche Zeitschrift
für Forst–und Land- wirtschft 11(5/6):300–355,
1913).
Hollinger, DY. 1987. Gas exchange and dry matter
allocation responses to elevation of atmospheric
CO2 concentration in seedlings of three tree species.
Tree Physiology 3:193–202.
Holm, F. 1940. Douglasgran, proveniens og vaekst
(with Ger- man summary: Die Douglasie,
Provenienz und Wachs- tum). Det forstlige
Forsøgsvaesen 15(4):233–312.
Holmes, GD. 1951. Methods of testing the germination
quality of forest tree seed, and interpretation of
results. Forestry Abstracts 13(1) 5–14.
Holmes, GD, and G Buszewicz. 1958. The storage of
seed of temperate forest tree species. Forestry
Abstracts 19:313–322, 455–476.
Hooker, WJ. 1838. Flora Boreali–Americana, or the
botany of the Northern parts of British America.
Adolphus Richter & Co., London 1834–1840. Vol.
2, pt. 7.
Hooven, EF. 1955. A Field Test of Tetramine-treated
Douglas-fir Seed. Research Note No. 21. Oregon
State Board of For- estry, Salem, OR. Online
http://hdl.handle.net/1957/8021
Hooven, EF. 1956. Field Test of Tetramine Treated
Douglas-fir Seed. Research Note No. 29, Oregon
State Board of Forestry, Salem, OR. 11 p. Online
http://hdl.handle.net/1957/8029
Hooven, E. 1958. Deer Mouse and Reforestation in the
Tillamook Burn. Oregon Forest Lands Research
Center. Forest Re- search Note No. 37, Corvallis,
OR. 31 p. Online http:// hdl.handle.net/1957/8037
Hopkins, DM, and DC Benninghoff. 1953. Evidence of a
very warm Pleistocence interglacial interval on
Seward Peninsula, Alaska. Geological Society of
America Bulletin 64(12 pt. 2):1435–1436.
Hosie, RC. 1969. Native trees of Canada. Canadian Forest Service,
Department of Fisheries and Forestry, Ottawa, Canada.
Howe, GT, K Jayawickrama, M Cherry, GR Johnson,
and NC Wheeler. 2006. Breeding Douglas-fir.
Plant Breeding Review 27:245–353.
Hulme, MA, and GE Miller. 1988. Potential for control
of Barbara colfaxiana (Kearfott) (Lepidoptera:
Olethreutidae) using Trichogramma spp., pp. 483–
488 in Proceedings of an International Symposium
on Trichogramma and Other Egg parasites, ed. J
Voegle, J Waage, and J van Lenteren. INRA
Publications, Versailles, France.
Hummel, FC, and J Christie. 1953. Revised Yield Tables
for Conifers in Great Britain. Forestry Commission,
Forestry Record No. 24.
Hummel, S. 2009. Branch and crown dimensions of
Douglas-fir trees harvested from old-growth forests
in Washington, Oregon and California during the
1960s. Northwest Sci- ence 83(3):239–252.
Hunt, GA, and R Fogel. 1983. Fungal hyphal dynamics
in a western Oregon Douglas-fir stand. Soil Biology
and Bio- chemistry 15(6):641–649.
Huss, J, and H Siebert. 1976. Erfahrungen mit der
Kultur der Douglasie. Ergebnisse einer Umfrage
in hessischen und bayerischen Forstämtern.
Allgemeine Forstzeitschrift 31(15):279–284.
Hussey, NW. 1954. Megastigmus flies attacking conifer
seed. Forestry Commission Leaflet 8, H.M.
Stationery Office, London.
Hussey, NW. 1955. The life-histories of Megastigmus
310 Douglas-fir: The Genus Pseudotsuga
London 106:133–151. Department of Agriculture, Washington, DC. 45 p.
Hussey, NW. 1956. The extent of seed loss in Douglas-fir Isaac, LA. 1943. Reproductive habits of Douglas-fir. Published
caused by Megastigmus. Scottish Forestry10(4):191– for the US Forest Service by the Charles Lathrop Pack
197.
Husted, L. 1991. Low Soil Temperature and Efficiency of
Ectomy- corrhizal Fungi. Unpublished PH D Thesis,
University of British Columbia, Vancouver, BC Canada,
226 p.
Husted, L, and DP Lavender. 1989. Effect of soil temperature
upon the root growth and mycorrhizal formation of
white spruce (Picea glauca (Moench Voss) seedlings
grown in controlled environments. Annales des Sciences
Forestières 46 (Supplement):750s-753s.
Hutchison, R. 1873. On the value of Corsican, Austrian,
and Douglas-firs as timber trees in Great Britain, and
on their adaptation to different soils and situations.
Royal Scottish Arboricultural Society Transactions
7:52–59.
Hutchison, R. 1879. On the progress of forestry in Scotland.
Royal Scottish Arboricultural Society Transactions 9:1–26.
Illingworth, K. 1978. Douglas-fir provenance trials in coastal
British Columbia: results to six years after planting.
pp. 411–425 in Proceedings of the IUFRO Joint Meeting
of Working Parties, Douglas-fir, Lodgepole Pine, Sitka
Spruce and True Firs, 21 Aug.–9 Sept. 1978. Ministry of
Forests, Vancouver, BC.
Illingworth, K, and D St. Pierre. 1975. Coastal Douglas-fir
provenance study. British Columbia Forest Service,
For- est Research Division, 5 unnumbered pages,
typescript.
Ilvessalo, L. 1926. Über die Anbaumöglichkeit ausländischer
Holzarten in spezieller Hinsicht auf die finnischen Ver-
hältnisse. Mitteilungen der Deutschen Dendrologischen
Ge- sellschaft 36:96–132.
Imbault N, Tardieu I, Joseph C, Zaerr JB, Bonnet-Masimbert
M. 1988. Possible role of isopentenyladenine and iso-
pentenyladenosine in flowering of Pseudotsuga
menziesii: endogenous variations and exogenous
applications. Plant Physiology and Biochemistry
26:289–295.
Instituto Forestal. 1986. Especies forestales exóticas de
interés económico para Chile, pp. 139–145 in Gerencia
de Desar- rollo, AF 86/32, Santiago, Chile, May 1986.
International Seed Testing Association. 1985. International
Rules for Seed Testing. Seed Science and Technology
13:2. 299–513.
Ionescu, A, and C Lazarescu. 1966. Duglasul, pinul strob si
stejarul rosu in culturile din Republica Socialista
Romania. Institutul de Cercetari Forestiere. Centrul de
Documentare Technica Pentru Economia Forestiera.
Bucharest.
Irgens-Møller, H. 1957. Ecotypic response to temperature
and photoperiod in Douglas-fir. Forest Science 3:79–83.
Irgens–Møller, H. 1963. Douglas-fir Archives and Their Use
in Tree Improvement. World Consultation on Forest
Genet- ics and Tree Improvement, Stockholm, Aug. 23–
30, 1963. FAO/FORGEN 63:3–8. FAO, Home.
Irvine, DR, DE Hibbs, and JPA Shatford. 2009. The relative
importance of biotic and abiotic controls on young
conifer growth after fire in the Klamath-Siskyou region.
Northwest Science 83(4):334–347.
Isaac, LA. 1927. Tree-Seed Flight Measured as Aid in
Reforesta- tion. pp. 647–649 in What’s New in Agriculture,
Yearbook of Agriculture. USDA, Washington, DC.
Isaac, LA. 1929. Flight of Douglas-fir seed. Forest Worker
5(5):16.
Isaac, LA. 1930. Seed flight in the Douglas-fir region.
Journal of Forestry 28:492.499.
Isaac, LA. 1935. Life of Douglas-fir seed in the in the forest
floor. Journal of Forestry 33:61–66.
Isaac, LA. 1938. Factors Affecting the Establishment of
Douglas- fir Seedlings. USDA Circular No. 486, US
Forestry Foundation, Washington, DC. 107 p. Wissenschaftliche Zeitschrift Technische Hochschule
Isaac, LA. 1949. Better Douglas-fir Forests From Better Dresden 6(8):543–548.
Seed. Col- lege of Forestry, University of
Washington Press, Seattle. Pacific Northwest Range
and Experiment station. USDA Forest Service. 65
p.
Isaac, LA. 1960. Problems and proposals for
international certification of tree seed origin and
stand quality with particular reference to western
North American species. pp. 690-696 in Fifth
World Forestry Congress Proceedings Vol. 2:
Genetics and Tree Improvement. University of Wash-
ington, Seattle.
Issac, LA, and EJ Dimock, II. 1958. Silvicultural
Characteristics of Douglas-fir. USDA Pacific
Northwest Forest and Range Experiment Station
Silvical Series No. 9. 18 p.
Isaac, LA, and EJ Dimock. 1960. Natural reproduction
of Doug- las-fir in the Pacific Northwest. Pacific
Northwest Forest and Range Experiment Station,
USDA Forest Service, Portland, OR.
Islam MA, Apostol KG, Jacobs DF, and Dumroese RK.
2008. Physiological responses of planting frozen
and thawed Douglas-fir seedlings, pp. 126–134 in
National Proceedings: Forest and Conservation
Nursery Associations—2007, coord. RK Dumroese
and LE Riley, Proceedings RMRS-P-57, USDA
Forest Service, Rocky Mountain Research Station,
Fort Collins, CO.
Ivanova, EV. 1963. Results of the introduction of
coniferous species into Belorussia, with English
summary: Concern- ing the results of the
introduction of conifers into the Byelorussian SSR,
pp. 149–158 in Proceedings of the First Scientific
Conference on the Research of Plant Resources and
their Enrichment in the Baltic Republics and
Byelorussia. The Academy of Sciences of the
Lithuanian SSR, Botanical Institute, Vilnius.
Jacobs, DF, Haase, DL, and Rose, R. 2005. Growth and
foliar nutrition of Douglas-fir seedlings provided
with supple- mental polymer-coated fertilizer.
Western Journal of Applied Forestry 20:58–63.
Jacobs, DF, R Rose, and DL Haase. 2003. Development
of Douglas-fir Seedling Root Architecture in
Response to Localized Nutrient Supply. Canadian
Journal of Forest Research 33(1):118–125.
Jacobs DF, R Rose, DL Haase, and PO Alzugaray. 2004.
Fer- tilization at planting impairs root system
development and drought avoidance of Douglas-fir
seedlings. Annals of Forest Science 61:643–651.
Jacobs, DF, AS Davis, BC Wilson, RK Dumroese, RC
Goodman, and KF Salifu. 2008. Short-day
treatment alters Douglas-fir seedling dehardening
and transplant root proliferation at varying
rhizosphere temperatures. Canadian Journal of
Forest Research 38:1526–1535.
Jackson, DJ, and GB Sweet. 1972. Flower initiation in
temperate woody plants. Horticultural Abstracts 42:9–
24
Jaffe, MJ, and S Forbes. 1993. Thigomorphogenesis:
The ef- fect of mechanical perturbation on plants.
Journal of Plant Growth Regulation 12:313–320.
Jäger, H, and L Beissner. 1884. Die Ziergehölze der Gärten
und Parkanlagen. BF Voigt, Weimar. 629 p.
Jahn, G. 1959. Über die standortbedingte
Höhenwuchsleis- tung der grünen Douglasie und
damit verbundenen Anbaufragen in verschiedenen
deutschen Wuchsräumen. Allgemeine
Forstzeitschrift 14(8):152–155.
Jahnel, H. 1959. Über Frostresestenz bei Waldbaumen.
Archiv für Forstwesen 8(8):697–725.
Jahnel, H. and Watzlawik, G. 1957. Mitteilungen aus
dem Forstbotanischen Garten Tharandt. II.
Fortführung der Beobachtungen Uber die Frosthärte
einiger Gehölze im Winter 1955/56.
References 311
Report PSW-GTR-143, Albany, CA. 219 p.
James, RL. 1986. Diseases on conifer seedlings caused by Jensen, LA, and E Noll. 1959. Experience in germination testing of
seed- borne Fusarium species. pp. 267–269 in Pacific Northwest Douglas-fir seed. Proceedings of the
Proceedings–Conifer Tree Seed in the Inland Mountain Association of Official Seed Analysts 49(1):107–113.
West. Symposium. comp. RC Scheanner. USDA Forest Jensen, M. 1996. Breaking of three seed dormancy at controlled
Service Information Research Station. General Technical
Report INT 203. 287 p.
James, RL. 1986. Occurrence of Fusarium in Douglas-fir Seed
and Containerized Seedlings of the Plum Creek Nursery,
Pablo, Montana. USDA Forest Service, Northern Region
Report 86–4. 10 p.
James, RL, and JY Woo. 1987. Pathogenicity of Alternaria
Al- ternata on Young Douglas-fir and Engelmann Spruce
Germ- lings. USDA Forest Service, Northern Region,
Forest Pest Management Report 87–9.
James, RL, RK Dumroese, and DL Wenny. 1997.
Pathogenicity of Fusarium proliferatum in container-
grown Douglas-fir seedlings, pp. 26–33 in Proceedings
of the 3rd meeting of IUFRO working party S7.03–04
Diseases and insects in forest nurseries, ed. RL James.
USDA Forest Service Northern Review Forest Health
Protection.
James, RN, and EH Bunn. 1978. Factors to be considered
in formulating silvicultural regimes for Douglas-fir,
pp. 341–343 in A Review of Douglas-fir in New
Zealand. FRI Symposium No. 15. Forest Research
Institute. Rotorua.
Jameson, Jr. EE. 1952. Food of deermice, Peromyscus
manipula- tus and Peromyscus boylei in the northern
Sierra Nevada, California. Journal of Mammalogy
33:50–60.
Janzen, DH. 1971. Seed predation by animals. Annual
Review of Ecology and Systematics 2(Nov):465–492. DOI:
10.1146/ annurev.es.02.110171.002341
Jankauskas, M. 1951. Prospects for the introduction of fast
growing conifers into the forests of Lithuania, pp. 228–
237 in Proceedings Science Session on problems in
biology and agriculture, Riga, Oct. 22–26, 1951.
Akademija Nauk SSSR.
Jaquish, BC. 1990. Geographic variation in ten–year height
growth of interior Douglas-fir in British Columbia. Paper
2.144 in Proceedings Joint Meet. WFGA and IUFRO WP’s
S2.02–05, 06, 12, 14. Olympia, WA. Weyerhaeuser Co.,
Centralia, WA.
Jarvis, SB, MA Taylor, and HV Davies. 1997. Molecular
Changes Associated with Dormancy-Breakage in Douglas-fir
Tree Seeds, pp. 225–260 in Basic and applied aspects of
seed biology, ed. RH Ellis, M Black, AJ Murdoch, and TD
Hong. Kluwer Academic Publishers, Dordrecht, Germany.
823 p.
Jarvis, SB, MA Taylor, MR Maclead, and HV Davies.
1996. Cloning and characterisation of the cDNA
clones of three genes that are differentially expressed
during dormancy- breakage in the seeds of Douglas-fir
(Pseudotsuga menzie- sii). Journal of Plant Physiology
147(5):559–566.
Jarvis, SB, MA Taylor, D Bianco, F Corbineau, and HV Da-
vies. 1997a. Dormancy-breakage in seeds of Douglas-fir
(Pseudotsuga menziesii (mirb) Franco). Support for the
hypothesis that LEA gene expression is essential for this
process. Journal of Plant Physiology 151(4):457–464.
Jarry, M, J-N Candau, A Roques, and B Ycart. 1997. Impact
of emigrating seed chalcid, Megastigmus spermotrophus
Wachtl (Hymenoptera: Torymidae), on seed production
in a Douglas-fir seed orchard in France and modelling of
orchard invasion. The Canadian Entomologist 129(1):7–
19.
Jenkinson, JL, and JA Nelson, JA. 1984. Cold Storage
Increases Dehydration Stress in Pacific Douglas-fir.
USDA Forest Service General Technical Report INT-
185, pp. 38–44.
Jenkinson, JL, JA Nelson, and ME Huddleson. 1993.
Improv- ing planting stock quality-the Humoldt
experience. USDA Forest Service General Technical
Forestry and Rural Development Publication 1168. 11
moisture content. Combined Proceedings p.
International Plant Propagators Society 46:296–304. Johnson, NE, and HJ Heikkenen. 1958. Damage to the seed
Jestaedt, M. 1980. Untersuchungen über die of Douglas-fir by the Douglas-fir cone midge. Forest
Jungendentwicklung von Douglasienprovenienzen in Sci- ence 4:274–282.
Hessen. JD Sauerländer’s Verlag, Frankfurt/Main. 105 Johnson, NE, and JH Rediske. 1964. Tests of systemic insec-
p.
Johnson, ALS. 1971. Residual effects of sunscald and
frost injury on young Douglas-fir. Bi-monthly
Research Notes 27(2):14–15. Department of
Fisheries and Forestry Canada.
Johnson, DW, and RD Harvey. 1975. Seed-protection
fun- gicides for control of Douglas-fir and
ponderosa pine seedling root rots. Tree Planters’
Notes 26 (2):3–5.
Johnson, DR, GP Markin, RC Reardan, and WK
Randalls. 1984. Injecting Metasystox-R® at three
spacing intervals to improve seed yield in Douglas
fir. Journal of Economic Entomology 77(5):1320–
1322.
Johnson, FD. 1961. Native trees of Idaho. University of Idaho,
Agricultural Extension Service Bulletin 289.
Johnson, GR. 1997. Site-to-site genetic correlations
and their implications on breeding zone size and
optimum num- ber of progeny test sites for
coastal Douglas-fir. Silvae Genetica 46:280–285.
Johnson, GR. 2002. Genetic variation in tolerance of
Douglas-fir to Swiss needle cast as assessed by
symptom expression. Silvae Genetica 51(2/3):80–
86.
Johnson, GR, RA Sniesko, and NL Mandel. 1997.
Age trends in Douglas-fir genetic parameters and
implications for optional selection age. Silvae
Genetica 46(6):349–358.
Johnson, JD. 1982. The effects of photoperiod during
cold storage on the survival and growth of loblolly
pine seed- lings. pp. 401–408 in Proceedings of the
2d Biennial Southern Silvicultural Research
Conference. U.S. Forest Serv., General Technical
Report SE.24. Southeastern Forest Exp. Sta.,
Asheville, NC.
Johnson, JD, and WK Ferrell. 1982. The relationship of
abscisic acid metabolism to stomatal conductance in
Douglas-fir during water stress. Physiologia
Plantarum 55(4):431–437.
Johnson, LC, and H Irgens-Møller. 1964. The effect of
photo- period and light quality upon germination of
Douglas-fir seed. Forest Science 10(2):200–205.
Johnson, NE. 1962a. Distribution of Douglas-fir cone
midges in the forest litter beneath young, open-
grown Douglas- fir. Canadian Entomologist
94:915–921.
Johnson, NE. 1962b. Tests of Guthion for the control of
the Douglas-fir cone midge. Journal of Economic
Entomology 55:615–616.
Johnson, NE. 1963a. Contarinia washingtonensis (Diptera:
Ceci- domyiidae), New species infesting the cones of
Douglas- fir. Annals of the Entomological Society of
America 56(1):94.
Johnson, NE. 1963b. Insecticides tested for control of
the Douglas-fir cone midge. Journal of Economic
Entomology 56:236–237.
Johnson, NE. 1963c. Helicopter application of Guthion
for control of the Douglas-fir cone midge. Journal
of Economic Entomology 56:600-603.
Johnson, NE. 1964. Chemical control of the Douglas-fir
cone midge Contarinia oregonensis using a
mistblower from a truck-mounted ladder. Journal
of Economic Entomology 57:556–558.
Johnson, NE, and SW Meso. 1966. Effectiveness of
three sys- temic insecticides for Douglas-fir cone
and seed insect control. Weyerhaeuser For. Pap. 10.
10 p.
Johnson, NE, and AF Hedlin. 1967. Douglas-fir
insects and their control. Canada, Department of
312 Douglas-fir: The Genus Pseudotsuga
ticides for the control of the Douglas-fir cone midge, Deutschen
Contarinia oregonensis Foote. Weyerhaeuser
Company Forestry Research Note 56. 13 pp.
Johnson, NE, and JH Rediske. 1965. A test of systemic
insecti- cides to control Douglas-fir cone and seed
insects. Journal of Economic Entomology 58:1020–
1021.
Johnson, NE, and JK Winjum. 1960. Douglas-fir cone and seed
insect biological and control studies: Progress in 1958,
1959. Forest Research Note 22, Weyerhaeuser Timber
Company, Tacoma, WA. 23 pp.
Johnson, NE, and JG Zingg. 1967. Effective translocation
of four systemic insecticides following application to
the foliage and cones of Douglas-fir. Journal of
Economic Entomology 60:575–578.
Johnson, R. 1998. Breeding design considerations for coastal
Douglas-fir. USDA Forest Service Pacific Northwest Re-
search Station. PNW-General Technical Report GTR-411,
Portland, OR. 34 p.
Johnson, R, and Temel, F. 1998. The genetics of Swiss
needle- cast intolerance, pp. 20–21 in Coast
Cooperative Annual Report, Forest Research
Laboratory, Oregon State Uni- versity, Corvallis.
Johnson, R, and F Temel, F. 1999. Genetics of Swiss
needlecast tolerance - early screening, and field results,
pp. 10–11 in Swiss Needle Cast Cooperative Annual
Report, 1999, ed.
G. Philip. Oregon State University, College of
Forestry, Forest Research Laboratory, Corvallis.
Johnson-Flanagan, AM, and JN Owens. 1985. Root Growth
and Root Growth Capacity of Whitespruce (Picea
glauca (Moench) Voss Seedlings. Canadian Journal of
Forest Re- search 15:625–630.
Joly, RJ, WT Adams, and SG Stafford. 1989. Phenological
and morphological responses of mesic and dry site
sources of coastal Douglas-fir to water deficit. Forest
Science 35(4):987–1005.
Jones, MM, NC Turner, and CB Osmund. 1981. Mechanisms
of drought resistance p. 15–37 in Physiology and
Biochemistry of Drought Resistance in Plants, Academic
Press Australia.
Jones, SK, and PG Gosling. 1994. “Target moisture
content” prechill overcomes the dormancy of
temperate conifer seeds. New Forests 8:309–321.
Joseph, G, and RG Kelsey. 1998. Ethanol synthesis,
nitrogen, carbohydrates, and growth in tissues from
nitrogen fer- tilized Pseudotsuga menziesii (Mirb.)
Franco and Pinus ponderosa Dougl. ex Laws. seedlings.
Trees 13:103–111.
Joseph, G, and RG Kelsey. 2000. Physiology and growth of
Douglas-fir seedlings treated with ethanol solutions.
Plant Science 150:191–199.
Joseph, G, and RG Kelsey. 2004. Ethanol synthesis and aero-
bic respiration in the laboratory by leader segments of
Douglas-fir seedlings from winter and spring. Journal of
Experimental Botany 55(399):1095–1103.
Junttila, O. 1988. To be or not to be dormant: Some com-
ments on the new dormancy nomenclature.
HortScience 23, 805–806.
Kahl, A. 1930. Der Winterfrost 1928/29 und seine
Auswirkun- gen auf Baum und Strauch Mitteilungen
Deutsche Den- drologische Gesellschaft 42:222–245.
Kalutski, KK, GV Krylov, and NA Bolotov. 1981.
Introduction of tree species for the forests of the future:
experience and prospects. Lesnoi Zhurnal 24(5):6–14.
Kangur, R. 1954. Shrews as Tree Seed Eaters in the Douglas-fir
Region. Research Note 17. Oregon State Board of Forestry,
Salem, OR.
Kangur, R. 1973. Snow damage to young western hemlock and
Douglas-fir. OSU Forest Research Laboratory, Corvallis,
Research Paper 21. 11 p.
Kanzow, H. 1936. Die Douglasie. Mitteilungen der
Dendrologischen Gesellschaft 48:235–245. Khan, SR, R Rose, DL Haase, and TE Sabin. 1996. Soil
Kanzow, H. 1937. Die Douglasie. Aufstellung einer water stress: Its effects on phenology, physiology, and
Ertragstafel auf Grund der Ergebnisse der mor- phology of containerized Douglas-fir seedlings.
Preussischen Probeflaechen and Auswertung von New Forests 12:19–39.
Provenienzversuchen. Zeitschrift für Forst– and Khan, SR, R Rose, DL Haase, and TE Sabin. 2000. Effects of
Jagdwesen 69:65–93, 113–139, 241–271. shade on morphology, chlorophyll concentration, and
Kaplan, R. 1937. Über die Bildung der Stele aus dem
Urmeri- steme von Pteridophyten und
Spermatophyten. Planta 27:224–268.
Kaspar, TC, and WL Bland. 1992. Soil temperature and root
growth. Soil Science 154(4):290–299.
Kay, J, and ML Anderson. 1928. Douglas-fir at home
and abroad. Empire Forestry Journal 7(1):22–40.
Kaya, Z, WT Adams, and RK Campbell. 1994.
Adaptive sig- nificance of intermittent shoot
growth in Douglas-fir seedlings. Tree Physiology
14:1277–1289.
Keen, FP. 1952. Insects affecting seed production, p. 17–
27. In Insect enemies of western forests. United States
Department of Agriculture Miscellaneous Publication
No. 273.
Keen, FP. 1953. Correction of identity of the Douglas-fir
cone
moth. Journal of Economic Entomology 46:1107–1108.
Keen, FP. 1958. Cone and Seed Insects of Western
Forest Trees, California Range and Experiment
Station, Forest Service Technical Bulletin 1169.
U.S. Department of Agriculture, Government
Printing Office, Washington, DC. 168 p.
Kelsey, RG, G Joseph, and E Gerson. 1998. Ethanol
synthesis, nitrogen, carbohydrates, growth in tissues
from nitrogen fertilized Pseudotsuga menziesii
(Mirb.) Franco and Pinus ponderosa Dougl. ex
Laws. seedlings. Trees 13(2):103–111.
Keltjens, WG. 1990. Effects of aluminum on growth
and nu- trient status of Douglas-fir seedlings
grown in culture solution. Tree Physiology
6:165–175.
Kenk, G, and J Hradetzky. 1984. Behandlung und
Wachstum der Douglasie in Baden–Württemberg.
Mitteilungen der Forstlichen Versuchs– und
Forschungsanstalt Baden–Würt- temberg, Heft 113.
Kenk, G, and M Thren. 1984a. Ergebnisse verschiedener
Douglasienprovenienzversuche in Baden–
Württemberg. Teil I. Der Internationale
Douglasien–Provenienzversuch. Allgemeine Forst-
und Jagdzeitung 155(7/8):165–184.
Kenk, G, and M Thren. 1984b. Ergebnisse verschiedener
Doug- lasienprovenienzversuche in Baden–
Württemberg. Teil II. Die Versuche Kirchzarten,
Aalen/Schwarzach, Steinheim and
Heidelberg/Ettenheim/ Kandern. Allgemeine Forst–
und Jagdzeitung 155(10/11):221–240.
Kermode, AR. 1995. Regulatory mechanisms in the
transition from seed development to germination:
interactions be- tween the embryo and the seed
environment, pp. 273–332 in Seed Development
and Germination. ed. N Kigel and G Galili.
Marcel Dekken, Inc. New York. 853 p.
Kershaw, JA, and DA Maguire. 1995. Crown structure
in Western hemlock, Douglas-fir, and grand fir in
Western Washington: trends in branch-level mass
and leaf area. Canadian Journal of Forest Research
25:1897–1912.
Kershaw, JA Jr, DA Maguire, and DW Hann. 1990.
Longevity and duration of radial growth in
Douglas-fir branches. Canadian Journal of Forest
Research 20:1690-1695.
Ketchie, DO, and W Lopushinsky. 1981. Composition
of Root Pressure Exudate From Conifers. Pacific
Northwest Forest and Range Experiment Station
Research Note, Portland, OR.
Keyes, J, and CF Smith. 1943. Pine Seed-Spot Protection
with
Screens in California. Journal of Forestry. 41:443–444.
References 313
Ergebnisse aus dem internationalen Douglasien–
chlorophyll fluorescence of four Pacific Northwest Herkunftsversuch von 1970 in der Bundesrepublik
conifer Deutschland. Silvae Genetica 28(5/6):226–244.
species. New Forests 19:171–186.
Killius, R. 1931. Anbauversuche fremdländischer Holzarten
in badischen Waldungen nach dem Stand von 1929/30.
MIt- teilungen aus dem Forstlichen Versuchswesen Badens,
Heft 3.
Kimming, JP. 1992. Balancing Act Environmental Issues in
For- estry. UBC Press, Vancouver, BC.
King, JH, FC Yeh, JC Heaman,and BP Dancik. 1988a. Selec-
tion of wood density and diameter in controlled crosses
of coastal Douglas-fir. Silvae Genetica 37(3/4):152–157.
King, JH, FC Yeh, and JC Heaman. 1988b. Selection of
growth and yield traits in controlled crosses of coastal
Douglas- fir. Silvae Genetica 37(3/4):158–164.
King, JE, DD Marshall, and JF Bell. 2002. Levels-of-Growing-
Stock Cooperative Study in Douglas-fir: Report no. 17: The
Skykomish Study, 1961–93; The Clemons Study, 1963–94.
USDA Forest Service Research Paper PNW-RP-548, Port-
land, OR.
Kirkland, A. 1969. Notes on the establishment and thinning
of old crop Douglas-fir in Kaingaroa Forest. New
Zealand Journal of Forestry 14(1):25–37.
Kirkland, A. 1971. Altitudinal limits, pp. 92–94 in The Role
of Exotic Genera Other Than Pinus in New Zealand.
FRI Symposium No. 10. Forest Service, New Zealand
Forest Research Institute, Rotorua.
Kirkwood, JE. 1922. Forest Distribution in the Northern Rocky
Mountains. University of Montana, Bulletin 247.
Kitzmiller, JH. 1990. Genetic variation and adaptability of
Douglas-fir in Northwestern California. Pap. 2.181 in
Proceedings, Joint Meet. WFGA and IUFRO WP’s S2.02–
05, 06, 12, 14, Olympia. WA. Weyerhaeuser Co.,
Centralia, WA.
Kleinschmit, J. 1973. Zur Herkunftsfrage bei der Douglasie.
Der Forst and Holzwirt 28(11):209–213.
Kleinschmit, J. 1978. Douglas-fir in Germany. pp. 317–333 in
Background papers and Douglas-fir provenances. Vol. l:
IUFRO Joint Meeting of Working Parties S2.02–05
Douglas-fir Prov- enances, 52.02–06 Lodgepole Pine
Provenances, 2.02–12 Sitka Spruce Provenances, 2.02–14
Abies Provenances. Vancouver, Canada 1978. British
Columbia Ministry of Forests, In- formation Services
Branch, Victoria, BC.
Kleinschmit, J. 1984. Neuere Ergebnisse der Douglasien-
provenienzforschung und -züchtung in der Bundesre-
publik Deutschland. Schweizer Zeitschrift für
Forstwesen 135(8):655–679.
Kleinschmit, J, J Racz, H Weisgerber, W Dietze, H
Dieterich, and R Dimpflmeier. 1974. Ergebnisse aus
dem internation- alen Douglasienherkunftsversuch von
1970 in der Bundes- republik Deutschland. Silvae
Genetica 23(6):167–176.
Kleinschmit, J, J Svolba, H Weissgerber, R Dimpflmeier, W
Ruetz, and T Widmaier. 1987. Results of the IUFRO
Doug- las-fir provenance experiment in the Federal
Republic of Germany at age 14 in Proceedings IUFRO
Working Party S2.02–05 Meeting on Breeding Strategy
of Douglas-fir as an Introduced Species. Vienna, Austria
June 10–14, 1985. FBVA Berichte No.21:67–84.
Schriftenreihe der Forstlichen Bundesversuchsanstalt
Vienna, Austria.
Kleinschmit, J, J Svolba, H Weissgerber, H-M Rau, R
Dimpfl- meier, W Ruetz, and A Franke. Widmaier.
1990. Results of the IUFRO Douglas-fir provenance
experiment in the Federal Republic of Germany at age
20, paper 2.195 in Proceedings of the Joint Meeting of
Western Forest Genetics Association and IUFRO
Working Parties, S2.02-05, 06, 12 and 14. Olympia, WA,
USA. 10 p.
Kleinschmit, J, J Svolba, H Weissgerber, M Jestaedt, R
Dimp- flmeier, W Ruetz, and H Dieterich. 1979.
University of British Columbia, Vancouver, BC. 179
Klepac, D. 1962. Prilog poznavanju rasta i prirasta p.
zelene duglazije i americkog borovka (with Engl. Kozak, A. 1964. Sequential sampling for improving cone
summary: A contribution to the knowledge of col- lection and studying damage by cone and seed
Douglas-fir and East- ern white pine increment). insects in Douglas fir. Forestry Chronicle 40:210-218.
Summarski list 86 (1/2):10–31. Kozak, A, O Sziklai, BG Griffith, and JHG Smith. 1963. Variation
Klumpp, RT. 1999. Untersuchungen zur Genökologie
der Douglasie (Pseudotsuga (Mirb.) Franco.
Doctoral disserta- tion Fakultat für Forstwissen-
schaften und Waldökologie der Georg-August
Universität Göttingen, Germany.
Klumpp, R, and P Gürth. 1988. Die Einbringung der
Douglasie im Forstbezirk Sulzburg unter Karl
Philipp (1897–1910). Allgemeine Forst- und
Jagdzeitung 159(1/2):12–19.
Knell, G. 1966. Vorkommen und Verbreitung der
Doug- lasie in der Bundesrepublik. Allgemeine
Forstzeitschrift 21(34):579–581.
Knox, RB, and MB Singh. 1987. New perspectives in pollen
biology and fertilization. Ann Bot 60 [Suppl 4]:15–37.
Koerber, TW. 1960. Insects destructive to the Douglas-
fir seed crop in California: A problem analysis.
United States Depart- ment of Agriculture Forest
Service, Pacific Southwest Forest and Range
Experiment Station Technical Paper 45, Berkeley,
CA. 36 p.
Koerber, TW. 1962. Douglas-fir cone and seed insect
research progress report, 1959. United States
Department of Ag- riculture Forest Service, Pacific
Southwest Forest and Range Experiment Station, 37
p.
Koerber, TW. 1963. Leptoglossus occidentalis (Hemiptera:
Core- idae), a newly discovered pest of coniferous
seed. Annals of the Entomological Society of America
56:229–234.
Koerber, TW, and GP Markin. 1984. Metasystox-R®
injections increase seed yield of Douglas-fir in
California, Oregon, and Washington, pp. 137–146 in
Proceedings of the IUFRO Cone and Seed Insects
Working Party Conference. Working Party S2.07–01.
July 31–August 6, 1983, Athens, GA, ed. HO Yates,
III. USDA Forest Service Southeastern Forest
Experiment Station, Asheville, NC. 214 p.
Koller, D. 1972. Environmental Control of Seed
Germination, pp. 1–101 in Seed Biology, Vol II, ed.
TT Kozlowski. Academic Press, New York, 447 p.
Koller, D, and A Hadas. 1982. Water relations in the
germina- tion of seeds, in Physiological Plant
Ecology. II Water Rela- tions and Carbon
Assimilation, ed. OL Lange, PS Noble, CB
Osmond, and H Ziegler. Encyclopedia of Plant
Physiology. New Series Vol. 12b. Springer-Verlag,
Berlin.
Kolotelo, D. 1997. Anatomy and Morphology of Conifer
Tree Seed. Forest Nursery Technical Servies 1. 60 p.
British Columbia Ministry of Forests Victoria, BC.
Konar, R, and Y Oberoi. 1969. Recent work on
reproductive structures of living conifers and taxads
– a review. Botani- cal Review 35: 89–116.
Kotze, H. 2002. A strategy for growth and yield research
in pine and eucalypt plantations in Komatiland
Forests in South Africa, pp. 75–84 in Modelling
Forest Systems Workshop on the interface between
reality, modelling and the parameter estimation
processes, Sesimbra, Portugal, 2–5 June 2002, ed. A.
Amaro, D. Reed, and P. Soares. CABI. DOI:
10.1079/9780851996936.0075
Kotze, H, and F Malan. 2009. Further Progress in the
Develop- ment of Prediction Models for Growth and
Wood Quality of Plantation-Grown Pinus patula
Sawtimber in South Africa, USDA General Technical
Report GTR, 791, pp. 113–124.
Kozak, A. 1963. Analysis of some factors associated
with distribution and intensity of attack by cone
and seed insects by Douglas-fir. PhD dissertation,
314 Douglas-fir: The Genus Pseudotsuga
in Cone and Seed Yield From Young Open-Grown Silviculturae, Brno. 36: 275–286.
Douglas-firs of the UBC Research Forest. Research Note Kristöfel, F. 2003. Uber Anbauversuche mit fremdländischen
#57, 8 p. Faculty of Forestry, UBC. Vancouver, BC. Baumarten in Österreich. BFW Berichte No. 131. 81 p.
Kozlowski, TT. 1971. Growth and Development of Trees, Vol.
II: Cambial Growth, Root Growth, and Reproductive
Growth, Academic Press, New York, pp. 393–433.
Kozlowski, TT. 1992. Carbohydrate sources and sinks in
woody plants. Botannical Review 58:107–222.
Kozlowski, TT. 1999. Soil Compaction and Growth of
Woody
Plants. Scandinavian Journal of Forest Research 14:596–619.
Kozlowski, TT, and CR Gunn. 1972. Importance and
character- istics of seeds. pp. 1–20 in Seed Biology, ed.
TT Kozlowski. Volume 1 Academic Press, New York.
416 p.
Kozlowski, TT, and T Keller. 1966. Food relations of woody
plants. Botanical Review 32:293–382.
Kozlowski, TT, and SG Pallardy. 1997. Physiology of Woody
Plants, Second Edition. Academic Press, San Diego, CA.
411 p.
Kozuharov, JN. 1964. Results of planting Douglas-fir in the
Sliven region. Gorsko Stopanstvo 20(10):6–10.
Krajina, VJ. 1969. Ecology of Forest Trees in British Columbia.
Ecol- ogy of Western North America 2(1):1–150.
Department of Botany, University of British Columbia,
Vancouver.
Krajina, VJ, K Klinka, and J Worrall. 1982. Douglas-fir, pp. 55–
57 in Distribution and Ecological Characteristica of Trees
and Shrubs of British Columbia. University of British Co-
lumbia, Faculty of Forestry, Vancouver.
Kramer, PJ, and TT Kozlowski. 1979. Physiology of Woody
Plants. Academic Press, New York. 826 p.
Krasowski, MJ, and JN Owens. 2000. Morphological and
physiological attributes of root systems and seedlings
growth in three different Picea glauca reforestations
stock. Canadian Journal of Forest Research 30:1669–
1681.
Krauch, H. 1956. Management of Douglas-fir timberland in
the Southwest. USDA Forest Service, Station Paper 21.
Rocky Mountain Forest and Range Experiment Station,
Fort Collins, CO. 59 p.
Krauss, H. 1955. Betrachtungen Über die grüne Douglasie
in Thüringen. Forst und Jagd 5(6):250-252.
Kremser, W. 1974. Säen und Pflanzen. Das langfristige,
regio- nale Waldbauprogramm der Niedersächsischen
Landes- forstverwaltung und ihre
Walderneuerungsplaene nach der Sturmkatastrophe
vom 13. November 1972. Neues Archiv für
Niedersachsen 23(3):256–286.
Kriek, W. 1974. Douglas–fit IUFRO provenances in the
Neth- erlands, 1966–67 series. Nederlands
Bosbouwtijdschrift 46(1):1–14.
Kriek, W. 1975. Douglas-fir IUFRO provenances in the
Neth- erlands, 1968–69 series. Nursery results.
Nederlands Bos- bouwtijdschrift 47(3):100–116.
Kriek, W. 1978. Further development of Douglas–fit prov-
enances in the Netherlands 1966/67 series, pp.241–255 in
Vol. 1: Background and Douglas-fir Provenances.
Proceedings of the IUFRO Joint Meeting of Working
Parties S2.02–05 Douglas-fir Provenances, S2.02–06
Lodgepole Pine Prov- enances, S2.02–12 Sitka Spruce
Provenances, S2.02–14 Abies Provenances, Vancouver,
Canada 1978. British Columbia Ministry of Forests,
Information Services Branch, Victoria.
Kriek, W. 1983. Results in Dutch Field Trials with Douglas-
fir Provenances and Progenies of the 1968/69 Series.
Rijksinsti- tuut voor Onderzoek in de Bos– en
Landschapsbouw “De Dorschkamp” Wageningen.
Rapport No. 343.
Krístek, J. 1967. Die verbreitung der Douglasien-samensch-
lupfwespe, Megastigmus spermotrophus Wachtl. (Hym.,
Torymidae). Acta Universitatus Agriculturae et
Schriftenreibe des Bundesamtes und Forschungszentrums closed– cone pine flora from travertine near Little Sur,
fur Wald. Vienna, Austria. California. Madroño 17(2):33–51.
Krömer, K. 1903. Wurzelhaut, Hypodermis und Langlet, O. 1971. Two hundred years genecology. Taxon
Endodermis der Angiospermenwurzel. Bibliotheca 20(5/6):653–722.
Botanica 12, Heft 59:1–151.
Krueger, KW. 1967. Foliar Mineral Content of Forest- and
Nursery- Grown Douglas-fir Seedlings. U.S. Forest
Service Research Paper PNW-45, Pacific Northwest
Forest & Range Experi- ment Station, Portland,
Oregon. 12 p.
Krueger, KW, and JM Trappe. 1967. Food reserves and
sea- sonal growth of Douglas-fir seedlings. Forest
Science 13(2):192–202.
Krutina, F. 1927. Die Douglasfichte in der Schweiz. Mitteilungen
Deutsche Dendrologische Gesellschaft 38:231–235.
Kurmann, MH. 1990. Exine ontogeny in conifers, pp.
157–172 in Microspores: Evolution and Ontogeny,
ed. S. Blackmore and R.B. Knox, Academic Press,
London.
Laas, E. 1967. Dendrologia. Kirjastus “Valgus,” Tallinn.
Lacaze, JF, and R Tomassone. 1967. Contribution a
l’étude de la variabilité du douglas (Pseudotsuga
menziesii Mirb.). Characteristiques juveniles.
Annals of Forest Science 24(1):85–106.
Ladefoged, K. 1939. Untersuchungen über die
Periodizität im Ausbruch und Längenwachstum
der Wurzeln bei einigen unserer gewöhnlichsten
Waldbäume. Det forstlige Forsøgsvæsen i
Danmark 16:1–256.
Lähde, E, M Werren, K Etholen, and V Silander. 1984.
Ulkom- aisten havupuulajien varttuneista
viljelmistä Suomessa. (with English summary:
Older forest trials of exotic coni- fer species in
Finland.) Communicationes Instituti Forestalis
Fenniae 125.
Lait, CG, SL Bates, JH Borden, and AR Kermode. 2001.
Protein reserve hydrolysis in Douglas-fir seeds
Pseudotsuga men- ziesii (Mirb.) Franco following
feeding by the western co- nifer seed bug. Seed
Science and Technology 29: 609–617.
Lambert, AB. 1803. A description of the genus Pinus:
illustrated with figures, directions relative to the
cultivation, and remarks on the uses of the several
species. J White, London.
Lambert, AB. 1832. A description of the genus Pinus:
illustrated with figures, directions relative to the
cultivation, and remarks on the uses of the several
species. 3rd ed. Vol 2. Messrs Wed- dell, London
(irregular pages).
Landis, TW, RW Tinus, SE McDonald, and JP Barnett.
1989. The Container Tree Nursery Manual, Volume
4 Seedling Nutrition and Irrigation. Agr. Handbook
674, USDA Forest Service, Washington DC. 119 p.
Lane, N. 2005. Power, Sex, Suicide, Mitochondria and the
Meaning of Life. Oxford University Press, Oxford.
354 pp.
Lang, A. 1965. Effects of some internal and external
condi- tions on seed germination, pp. 848–893 in
Handbuch der Pflanzenphysiologie, ed. W Ruhland.
Springer, Berlin- Heidelberg-New York.
Lang, A. 1967. Plant growth regulation. Science. 157:589.
Lang, GA. 1994. Dormancy—the missing links:
molecular studies and integration of regulatory
plant and environ- mental interactions.
HortScience 29, 1255–1263.
Lang, GA, JD Early, NJ Arroyave, RL Darnell, GC
Martin, and GW Stutte. 1985. Dormancy: Toward a
reduced, universal terminology. HortScience
20(5):809–812.
Lang, GA, JD Early, GC Martin and RL Darnell. 1987. Endo-
, para-, and ecodormancy: physiological
terminology and classification for dormancy
research. HortScience 22:371–377.
Langenheim, JH, and JW Durham. 1963. Quarternary
References 315
Germination. Research Note No. 34, Oregon Forest Lands
Lanner, RM. 1966. Needed: a new approach to the study Research Center, Corvallis, OR. 15 p.
of Lavender, DP. 1958e. Effect of Ground Cover on Seedling Germi-
pollen dispersion. Silvae Genetica 15:50-52.
Lanner, RM. 1961. Living stumps in the Sierra Nevada.
Ecol- ogy 42:170–173.
Lanner, RM. 1963. Growth and Cone Production of Knob Cone
Pine Under Interrupted Nights. USDA Forest Service
Pacific Southwest Forest & Range Experiment Station
Research Note PSW-38.
Larsen, JB. 1976. Untersuchungen uber die
Frostempfindlich- keit von Douglasienherkünften and
über den Einfluss der Nährstoffversorgung auf die
Frostresistenz der Douglasie. Der Forst- und Holzwirt
31(15):299–302.
Larsen, JB. 1978a. Die Frostresistenz von 60
verschiedenen Douglasienherkünften sowie über den
Einfluss der Nähr- stoffversorgung auf die
Frostresistenz der Douglasie. Schriftenreihe aus der
Forstlichen Fakultät der Universitat Göttingen, Band 52,
126 p.
Larsen, JB. 1978b. Die Frostresistenz der Douglasie
Pseudotsuga menziesii (Mirb.) Franco). Verschiedener
Herkünfte mit unterschiedlichen Höhenlagen. Silvae
Genetica 27(3/4):150- 156.
Larsen, JB. 1980. Geographic variation in winter drought
resistance of Douglas-fir (Pseudotsuga menziesii
(Mirb) Franco) Silvae Genetica.
Larsen, JB. 1981. Geographic variation in winter drought
resistance of Douglas-fir (Pseudotsuga menziesii (Mirb.)
Franco). Silvae Genetica 30(4/5):109–114.
Larsen, JB. 1983. Trockenresistenz, Wasserhaushalt und
Wachstum junger Douglasien (Pseudotsuga menziesii)
und Küstentanne (Abies grandis) in Abhängigkeit von
der Nährstoffversorgung. Det forstlige Forsøgsvaesen
i Danmark 39:1–82.
Larsen, JB, and WF Ruetz. 1980. Frostresistenz
verschiedener Herkünfte der Douglasie (Pseudotsuga
menziesii) und der Küstentanne (Abies grandis) entlang
des 44. Breitengrades in Mittel Oregon.
Forstwissenschaftliches Centralblatt 99(4):222–233.
Larsen, JB, and HK Kromann. 1983. Provenienser of Doug-
lasgran (Pseudotsuga menziesii (Mirb.) Franco) i
Danmark. Det forstlige Forsøgsvaesen i Danmark
38(4):349–375.
Larsen, SC. 1940. Planting af Douglas. Dansk Skovforenings
Tidskrift 25(11):579–582.
Larson, PR. 1961. Influence of date of flushing on
flowering
in Pinus banksiana. Nature 192:82–83.
Latham, P, and JC Tappeiner. 2002. Response of old-growth
conifers to reduction in stand density in western Oregon
forests. Tree Physiology 22:137–146.
Lathrop, JK, and RA Mecklenburg. 1971. Root regeneration
and root dormancy in Taxus spp. Journal of the
American Society for Horticultural Science 96:111–114.
Laude, HH. 1939. Diurnal cycle of heat resistance in
plants.
Science 89:556–557.
Lavender, DP. 1954. Douglas-fir Seed Germination After
Exposure to Varying Temperatures and Humidities.
Research Note No. 18. Oregon State Board of Forestry,
Salem, OR. 3 p.
Lavender, DP. 1958a. Effect of Field Stratification on Douglas-
fir Seed Germination. Research Note 33, Oregon Forest
Lands Research Center, Corvallis, OR. 8 p.
Lavender, DP. 1958b. Viability of Douglas-fir Seed After
Storage in the Cones. Research Note No. 31, Oregon
Forest Lands Research Center, Corvallis, OR.
Lavender, DP. 1958c. Effects of Seed Size on Douglas-fir
Seedlings. Research Note No. 32, Oregon Forest Lands
Research Center, Corvallis, OR.
Lavender, DP. 1958d. Seeding Dates and Douglas-fir
Lavender, DP, and WS Overton. 1972. Thermoperiods and
nation and Survival. Research Note 38, Oregon Soil Temperatures as They Affect Growth and Dormancy of
State Uni- versity, Forest Research Laboratory, Douglas-fir Seedlings of Different Geographic Origin.
Corvallis, OR. 32 p. Forest Research Laboratory Research Paper 13, School of
For- estry, Oregon State University, Corvallis. 26 p.
Lavender, DP. 1962. The growth of some coniferous
species in a controlled environment. Doctoral Lavender, DP, and SN Silim. 1987. The role of plant growth
dissertation, Oregon State University, Corvallis.
Lavender, DP. 1964. Date of Lifting for Survival of
Douglas-fir Seedlings. Research Note 49, Oregon
State University, Forest Research Laboratory,
Corvallis, OR.
Lavender, DP. 1970. Foliar analysis and how it is used:
A Review. Forest Research Laboratory, School of
Forestry, Oregon State University, Corvallis. 8 p.
Lavender, DP. 1978a. Seeds, pp. 48–62 in Regenerating
Oregon’s Forest, ed. BD Eleany, RP Graeves, and
RK Hermann. Oregon State University Extension
Service, Corvallis, OR.
Lavender, DP. 1978b. Bud activity of Douglas-fir
seedlings receiving different photoperiods in cold
storage, pp. 245–248 in Proceedings of the Fifth
North American Forest Biology Workshop, March
13–15, 1978, ed. CA Hollis and AE Squillace,
School of Forest Resources and Conserva- tion,
University of Florida. 430 p.
Lavender, DP. 1981. Environment and Shoot Growth of
Woody Plants. Research Paper 45, Forest Research
Laboratory, School of Forestry, Oregon State
University, Corvallis.
Lavender, DP. 1984. Plant physiology and nursery
environ- ments: Interactions affecting seedling
growth, pp. 133–141 in Forest Nursery Manual:
Production of Bare-Rooted Seed- lings, ed. ML
Duryea and TD Landis. Forest Research
Laboratory, OSU, Corvallis, OR/ Martinus
Nijhoff/Dr.
W. Junk Publishers, Boston, MA. 385 p.
Lavender, DP. 1985. Bud dormancy, pp. 7–15 in
Proceed- ings: Evaluating seedling quality:
principles, procedures and predictive abilities of
major tests. Workshop held October 16–18, 1984.
Forest Research Laboratory, Oregon State
University, Corvallis.
Lavender DP. 1988. Characterization and manipulation
of the physiological quality of planting stock, in
Proceedings of the Tenth North. American Forest
Biology Workshop, ed. DL Lester and JG Worrall,
University of British Columbia, Vancouver.
Lavender, DP. 1989. Predicted global warming and the
chill- ing requirements of conifers, in Climate
Change in British Columbia: Implications for the
Forest Sector. FRDA Report 075, ed. DL
Spittlehouse and DFW Pollard, Forestry Canada,
Victoria, BC.
Lavender DP. 1991. Measuring phenology and
dormancy, pp. 403–422 in Techniques and
Approaches in Forest Tree Ecophysiology, ed. JP
Lassoie and TM Hinckley. CRC Press, Inc., Boca
Ratón, FL.
Lavender, DP, and RL Carmichael. 1967. Effect of three
vari- ables on mineral concentrations in Douglas fir
needles. Forest Science 12(4):441–446.
Lavender, DP, and BD Cleary. 1974. Coniferous
seedling pro- duction techniques to improve
seedling establishment, pp. 177–180 In
Proceedings of North American Container- ized
Forest Tree Seedlings Symposium, ed. RW Tinus,
WI Stein, and WE Balmer, Great Plains
Agricultural Council Publication 68.
Lavender, DP, and WH Engstrom. 1956. Viability of Seeds
From Squirrel-Cut Douglas-fir Cones. Research Note
27, Oregon State Board of Forestry, Salem, OR. 19 p.
Lavender, DP, and RK Hermann. 1970. Regulation of
the growth potential of Douglas-fir seedlings
during dor- mancy. New Phytologist 69(3):675–
694.
316 Douglas-fir: The Genus Pseudotsuga
regulators in dormancy in forest trees. Journal of Plant ton scleroxylon K. Schum. Commonwealth Forestry Review
Growth Regulation 6(1–2):171–191. 60(2):117–126.
Lavender, DP, and SN Silim. 1992. The relationship of pre-
harvest and storage photoperiod to the vigour of
seedlings of some boreal coniferous species, in
Proceedings of the 1991 Ontario Tree Seedling Growers
Association Workshop, Kirkland Lake, ON, ed. J.
Gillham, Ontario Ministry of Natural Resources,
Timmins, ON.
Lavender, DP, and SG Stafford. 1985. Douglas-fir seedlings:
some factors affecting chilling requirement, bud activity,
and new foliage production. Canadian Journal of Forest
Research 15(2):309–312.
Lavender, DP, and RB Walker. 1979. Nitrogen and related
elements in nutrition of forest trees, in Proc Forest
Fertil- ization Conference, ed. SP Gessel, RM Kenady,
and WA Atkinson. Institute of Forest Resources
Contribution 40. University of Washington, Seattle.
Lavender, DP, and PF Wareing. 1972. Effects of day
length and chilling on the responses of Douglas-fir
(Pseudotsuga menziesii (Mirb.) Franco) seedlings to
root damage and storage. New Phytologist 71:1055–
1067.
Lavender, DP, and DB Zaerr. 1985. Temperate forest trees,
pp. 1–12 in CRC Handbook of Flowering Vol. 1. CRC
press, ed. HH Halevy, Boca Raton, FL. 568 p.
Lavender, DP, MH Bergman, and LD Calvin. 1956. Natural
Regeneration on Staggered Settings. Research Bulletin No.
10. Oregon State Board of Forestry, Salem, OR. 36 p.
Lavender, DP, KK Ching, and RK Hermann. 1968. Effects of
en-
vironment on the development of dormancy and
growth
of Douglas-fir seedlings. Botanical Gazette 129(1):70–83.
Lavender, DP, RK Hermann, and JB Zaerr. 1970. Growth
potential of Douglas-fir seedlings during dormancy, pp.
209–220 in Physiology of Tree Crops, ed. LC Luckwill
and CV Cutting, Academic Press, New York. 380 p.
Lavender, DP, WP Nagel, and A Doerkson. 1967. Eriophyid
mite damage on Douglas-fir seedlings. Journal of
Economic Entomology 60(2):621–622.
Lavender, DP, GB Sweet, JB Zaerr, and RK Hermann. 1973.
Spring shoot growth in Douglas-fir may be initiated by
gibberellins exported from the roots. Science 182:838–
839.
Lawrence, WH, and DH Rediske. 1960. Radio-tracer
technique for determining the fate of broadcast.
Douglas-fir seed. Society of American Foresters
Proceedings. 1959: 99–101.
Lawrence, WH, and JH Rediske. 1962. Fate of sown
Douglas- fir seed. Forest Science 8(3):210–218.
Lawson, AA. 1909. The gametophytes and embryo of Pseu-
dotsuga douglasii. Annals of Botany 23(90/April):163–180.
Lazarescu, C, and A Lonescu. 1964. Cultura Duglasului
Verde Si a Pinului Strob. Editura Agro-Silvica,
Bucharest.
Leadem, CL. 1988. Dormancy and Vigour of Tree Seeds.
British Columbia Ministry of Forests, Forestry Research
Labora- tory, Victoria.
Leadem, CL. 1987. The role of plant growth regulators in
the germination of forest tree seeds in Hormonal Control
of Tree Growth, ed. SV Kossuth and SD Ross. Martinos,
Nijholf, Boston, MA. 237 p.
Leadem, C. 1996. A Guide to the Biology of Forest Tree Seed.
British Columbia Ministry of Forests, Victoria, BC.
Leadem, CL, RD Eremko, and IH Davis. 1990. Seed Biology
Collection and Post-harvest handling, pp. 193–205 in
Regenerating British Columbia’s Forests, ed. R Parish,
CM Johnson, G Montgomery, A Vyse, RA Willis, D
Winston, and DP Lavender. University of British
Columbia, Van- couver. 372 p.
Leakey, RRB, NR Ferguson, and KA Longman. 1981. Preco-
cious flowering and reproductive biology of Triplochi-
Lebrun, C. 1970. Prétraitement des graines de Douglas à 78:366–371.
l'eau oxygénée. Revue forestière française 22(4):473- Linderman, RG. 2000. Effects of mycorrhizas on plant
476 (In French with English summary). tolerance to diseases, pp. 345–366 in Arbuscular
[Pretreatment of Douglas seeds with oxygenated Mycorrhizas: Physiol-
water.]
Ledgard, NJ, and MC Belton. 1985. Diversification and
op- portunities for forestry in the South Island high
country. New Zealand Forestry 30(1):133–143.
Legat, CE. 1932. The cultivation of exotic conifers in
South Africa, pp. 285–308 in Conifers in
Cultivation. The Royal Horticultural Society,
London.
Leininger, TD, PR Miller, SL Schilling, and PH
Dunn, 1991. Seedling responses of western
conifers to simulated ambi- ent sulfur dioxide
exposures. Forest Science 37(6):1538– 1549.
Lemoine, K, and H Wirten. 1988. Douglasgran i
Sverige– förekomst och produktion. Sveriges
Lantbruksuniversitet, Institutionen för Skogsskötsel,
Umeå. Examensarbete no. 1.
Lester, DT. 1967. Variation in cone production of red pine in
relation to weather. Canadian Journal of Botany 45:1683–91.
Leutham, P, and D Tappeiner. 2002. Response of old-
growth conifers to reduction in stand density in
western Oregon forests. Tree Physiology 22:137–
146.
Lewis, JD, RB McKane, DT Tingey, and PA
Beedlow. 2000. Vertical gradients in
photosynthetic light response within an old-
growth Douglas-fir and western hemlock canopy.
Tree Physiology 20:447–456.
Li, HL. 1950. The coniferales of Taiwan. Taiwania 1:285–310.
Li, HL. 1975. Pteridiphyta and Gymnospermae. Vol. 1.
Flora of Taiwan. Epoch, Taipei, Taiwan.
Li, P, and WT Adams. 1989. Range-wide patterns of
allozyme variation in Douglas-fir (Pseudotsuga
menziesii). Canadian Journal of Forest Research
19(2):149–161.
Li, P, and WT Adams. 1993. Genetic control of bud
phenology in pole-size trees and seedlings of
coastal Douglas-fir. Canadian Journal of Forest
Research 23(6):1043–1051.
Li, P, and WT Adams. 1994. Genetic variation in
cambric phenology of coastal Douglas-fir.
Canadian Journal of Forest Research 24:1864–
1870.
Li, XJ, PJ Burton, and CL Leadem. 1994. Interactive
effects of light and stratification on the germination
of some British Columbia conifers. Canadian
Journal of Botany 72:1635–1646.
Lieffers, VJ, and RL Rothwell. 1986. Effects of depth of
water table and substrate temperature on root and
top growth of Picea mariana and Larix lancina
seedlings. Canadian Journal of Forest Research
16(6):1201–1206. DOI 10.1139/x86-214.
Liese, J. 1932. Zur Biologie der Douglasienschütte.
Zeitschrift für Forst- und Jagdwesen 64:680–03.
Liese, J. 1935. Die Anfälligkeit der Douglasien
gegenüber der Douglasienschütte. Der Deutsche
Forstwirt 17:959–761, 773–975.
Liese, J. 1936. Die Douglasienrassen and ihre
Anfälligkeit gegenüber der Douglasiennadelschütte
(Rhabdocline pseudotsugae). Mitteilungen der
Deutschen Dendrologischen Gesellschaft 48:25–26.
Lin, WF, PC Tsoong, and LM Chang. 1953. A Synopsis
of Trees and Shrubs in Taiwan. Taiwan Forestry
Research Institute, Taipei, Taiwan.
Linderman, RG. 1985. Microbial interactions in the mycor-
rhizosphere. pp. 117 120. In Proceedings of the 6th
North American conference on Mycorrhizae, June 25
29, 1984, Bend, Oregon, U.S.A. 471 pp.
Linderman, RG. 1988. Mycorrhizal interactions with
the rhizosphere microflora: The
mycorrhizosphere effect. Phytopathology
References 317
Liu, T. 1966. Study on the Phytogeography of the Conifers and
ogy and Function, ed. Y Kapulnik and DD Douds, Taxads of Taiwan. Taiwan Forestry Research Institute Bul- letin
Kluwer Academic Publishers, Dordrecht, Boston, No. 122.
London. 372 p.
Linderman, RG. 2008. The mycorrhizosphere phenomenon,
pp. 341–355 in Mycorrhiza Works, ed. F Feldman, Y
Kapulnik, and J Barr, Deutsche Phytomedizinische
Gesellschaft, Braunschweig, Germany.
Linderman, RG, and M Hoefnagels. 1993. Controlling root
pathogens with mycorrhizal fungi and beneficial
bacteria, pp. 132–135 in Proc. Western Forest Nursery
Association Meeting, 14–18 Sept. 1992, Fallen Leaf
Lake, CA. T Landis, tech. coord. USDA General
Technical Report RM - 221. Rocky Mountain Forest and
Range Experiment Station, Fort Collins, CO.
Linderman, RG, and FL Pfleger. 1994. General summary,
pp. 337–344 in Mycorrhizae and Plant Health, ed. FL
Pfleger and RG Linderman, The American
Phytopathological Society, St. Paul, Minnesota. 344
pp.
Lines, R. 1956. Provenance: Douglas-fir, p. 36 in Forestry
Com- mission, Report on Forest Research 1955. HMSO,
London.
Lines, R. 1957. Provenance studies. Douglas-fir, p. 40 in For-
estry Commission, Report on Forest Research 1956.
HMSO, London.
Lines, R. 1977. Seed origin: Douglas-fir, p. 17 in Forestry Com-
mission, Report on Forest Research 1977. HMSO, London.
Forestry Commission Bulletin, UK 1987
Lines, R. 1980. The IUFRO experiments with Douglas-fir in
Scotland, pp.297–303 in Background papers and Douglas-
fir provenances. Vol.1: IUFRO joint meeting of Working
Parties S2.02–05 Douglas-fir provenances, S2.02–06
Lodgepole pine provenances, S2.02–12 Sitka spruce
provenances, S2.02–14 Abies provenances. Vancouver,
Canada 1978. BC Ministry of Forests, Information
Services Branch, Victoria.
Lines, R. 1987. Choice of seed origins for the main forest
species in Britain. Forestry Commission Bulletin 66,
London, HMSO.
Lines, R, and AF Mitchell. 1967. Provenance and Douglas-
fir. pp. 64-69 in Forestry Commission, Report on Forest
Research 1967. Forest Comm, London. 194 p.
Lines, R, and AF Mitchell. 1968. Provenance: Douglas-fir,
pp. 71–73 in Forestry Commission, Report on Forest
Research 1968. HMSO, London.
Lines, R, and CJA Samuel. 1987. Results of the IUFRO
Doug- las-fir experiments in Britain at age 10 years,
pp. 31–48 in FBVA–Berichte No.21. Schriftenreihe
der Forstlichen Bundesversuchsanstalt, Vienna,
Austria.
Lines, R, AF Mitchell, and AJ Low. 1967. Provenance of
seed. Douglas-fir, p. 44 in For. Comm. Rep. For. Res.
for Year Ending March 1966. HMSO, London.
Lipa, OL. 1940. Supplement to the gymnosperms of gardens
and parks in the Ukrainian SSR. Botanicheskii Zhurnal
(Journal Botanique de l’Academie des Sciences de la
RSS d’Ukraine) 1(1):119–126.
Lipow, SR, Johnson, GR, St. Clair, JB, and KJS
Jajawickrama. 2003. The role of tree improvement
programs for ex situ gene conservation of coastal
Douglas-fir in the Pacific Northwest. Forest Genetics
10(2):111–120.
Little, EL, Jr. 1956. Pinaceae, Rutaceae, Meliaceae,
Anacardia- ceae, Tamacaricaceae, Cornaceae, Oleaceae,
Bignoniaceae of Nevada in Contributions toward a
Flora of Nevada 40. Beltsville, Maryland.
Little, EL, Jr. 1971. Atlas of United States Trees. Vol. 1.
Conifers and important hardwoods. USDA, Forest Service,
Miscel- laneous Publications 1146, Washington DC.
Little, EL, Jr. 1979. Checklist of United States Trees (Native
and Naturalized). USDA Forest Service, Agriculture Hand-
book No. 541.
Douglas-fir seedlings. Forest Science 30(3):628–634.
Livingston, GK. 1964. The Influence of Moisture and Lopushinsky, W, and TA Max. 1990. Effect of soil tempera-
Temperature on the Preservation of Douglas-fir
Pollen by Freeze-drying. M. Sc. thesis, Oregon State
University. Corvallis.
Livingston, GK, and KK Ching. 1967. The longevity
and fer- tility of freeze-dried Douglas-fir pollen.
Silvae Genetica 16:98–101.
Livingston, NJ. 1986. Water Relations, Survival and
Growth of Conifer Seedlings Planted on a High
Elevation South- facing Clearcut. Doctoral
dissertation, University of Brit- ish Columbia,
Vancouver. 154 p.
Locke, GML. 1987. Census of Woodlands and Trees 1979–82.
Forestry Commission Bulletin 63. London, H.M.S.O.
Lockhart, JA, and J Bonner. 1957. Effects of Gibberellic
Acid on the Photoperiod-Controlled Growth of
Woody plants. Plant Physiology 32:492—494
Loescher, WH, T McCaman, and JD Keller. 1990.
Carbohydrate reserves, translocation, and storage in
woody plant roots. HortScience 25(3):274–281.
Longman, KA. 1961. Factors affecting flower initiation in
cer- tain confiers. Proceedings of the Linnean Society
of London, 172:124–127.
Longman, KA. 1970. Initiation of flowering on first year
cuttings of Metasequoia glyptostroboides Hu and
Cheng. Nature 227:299–300.
Longman, KA. 1971. Possibilities of controlled
reproduction in trees, pp. 101–104 in Forest and
Woodland Ecology, ITE Symposium No. 8, ed. FT
Last and AS Gardiner. Institute of Terrestrial
Ecology, Cambridge, England.
Longman, KA. 1976. Some experimental approaches
to the problem of phase-change in forest trees.
Acta Horticul- turae 56:81–90.
Longman, KA. 1982. Effects of gibberellin, clone and
environ- ment on cone initiation, shoot growth and
branching in Pinus contorta. Annals of Botany 50
(2):247–257.
Longman, KA. 1985. Tropical forest trees, pp. 23–39 in
CRC Handbook of Flowering, Vol. 1, ed. AH
Halevy, CRC Press, Boca Raton, FL.
Longman, KA, MP Coutts, and MR Bowen. 1972. Tree
physi- ology, pp. 83–87 in Forestry Commission,
Report on Forest Research. HMSO, London.
Longman, KA, TAA Nasr, and PF Wareing. 1965.
Gravimor- phism in trees. The effect of gravity on
flowering. Annals of Botany 29(3): 459–473.
Look, W, Sutherland, JR, and LJ Slugget. 1975.
Fungicide treat- ment of seeds for dumping-off
control in British Columbia forest nurseries. Tree
Planters’ Notes 26(3):16–18, 28.
Loopstra, CA. 1984. Patterns of Genetic Variation Within
and Among Breeding Zones of Douglas-fir in
Southwestern Oregon. MS thesis, Oregon State
University, Corvallis.
Loopstra, CA, and WT Adams. 1989. Patterns of
variation in first-year seedling traits within and
among Douglas-fir breeding zones in Southwest
Oregon. Silvae Genetica 38 (5/6):235–243.
Loopstra, CA, and WT Adams. 1989. Patterns of
variation in first–year seedling traits within and
among Douglas-fir breeding zones in Southwest
Oregon. Silvae Genetica 38(5/6):235–243.
Lopez, HJ. 1973. Reforestation in Chile related to seed
supply problems, in IUFRO Working Party S2.01–06
Seed Problems, Proceedings of International
Symposium on Seed Processing, Bergen, Norway,
1973. Vol. II, Paper No. 18.
Lopushinsky, W. 1980. Occurrence of root pressure
exuda- tion in Pacific Northwest conifer seedlings.
For. Sci. 26(2), 275–279.
Lopushinsky, W, and MR Kaufmann. 1984. Effects of
cold soil on water relations and spring growth of
318 Douglas-fir: The Genus Pseudotsuga
ture on root and shoot growth and of bud burst Lyr, H, and G Hoffmann. 1967. Growth Rates and Growth
timing Periodicity of Tree Roots, pp. 181–236 in International
in conifer seedling transplants. New Forests 4:107–124.
Lorey, T. 1890. Anbauversuche mit fremdländischen
Holzarten in den Staatswaldungen Württembergs.
Allgemeine Forst– und Jagdzeitung 66(7):255–258.
Lorey, T. 1897. Anbauversuche mit fremdländischen
Holzarten in den Staatswaldungen Württembergs.
Allgemeine Forst– und Jagdzeitung 73:14–19, 83–87.
Lori, G, and MI Salerno. 2002. Fusaria population associated
with Douglas-fir and Ponderosa pine seeds in Argentina.
Seed Science & Technology 30:559–566.
Loudon, JC. 1838. Arboretum et fruticetum britannicum; or,
The trees and shrubs of Britain, native and foreign, hardy
and half-hardy, pictorially and botanically delineated, and
scientifically and popularly described; with their propagation,
culture, management, and uses in the arts, in useful and orna-
mental plantations, and in landscape-gardening; preceded by
a historical and geographical outline of the trees and shrubs
of temperate climates throughout the world. Printed for the
author, London.
Louis, J. 1935. L’ontogénèse du système conducteur dans
la pousse feuillée des Dicotylées et des
Gymnospermes. La Cellule 44:85–172.
Louro, V, and P Cabrita. 1989. Pseudotsuga–Contribuição para
o conhecimento da sua cultura em Portugal. Direcção–
Geral dos Servicos Florestais e Aquícolas, Lisbon,
Portugal. Estudos e Informação No. 298.
Lowe, WJ, and NC Wheeler. 1993. Pollen contamination in
seed orchards, pp. 49–54 in Advances in Pollen
Management.
U.S. Dept. of Agriculture, Forest Service, Agriculture
Handbook 698.
Lowell, EC, SA Willits, and RL Krahmer. 1992.
Deterioration of fire-killed and fire-damaged timber in
the western United States. USDA, For. Serv. Pacific
Northw. For. Res. Sta. General Technical Report PNW-
GTR-292, 27 p.
Lowry, WP. 1966. Apparent meteorological requirements
for abundant cone crop in Douglas-fir. Forest Science
12:185–192.
Lowry, WP. 1967. Biometeorological Inference and Plant
Environment Intercalations Ground Level Climatology,
pp. 3–14, in Biometeorology: Proceedings of the twenty-
eighth annual biology colloquim, 1967, ed. WP Lowry,
Oregon State University, Corvallis.
Lowry, GL, and CT Youngberg. 1955. The effect of certain
Site and soil factors on establishment of Douglas-fir on
the Tillamook burn. Proceedings - Soil Science Society
of America 19:378–380
Luis, JFS. 1989. Crescimento e competição em povoamentos
de Pseudotsuga menziesii (Mirb.) Franco em Portugal.
Docto- ral dissertation. Universidade de Trás-os-Montes
e Alto Douro, Vila Real, Portugal.
Lukin, AV. 1966. Natural regeneration of species of conifers
introduced into the Lipetsk region. Byulleten’ Glavnogo
Botanicheskogo Sada No. 61:11–13. Original not seen,
cited from Forestry Abstracts 28:2062, 1967.
Lundberg, J. 1957. Proveniensforsøg med douglasgran. Det
forstlige Forsøgsvaesen i Danmark 23:345–370.
Lundegardh, H. 1914. Das Wachstum des
Vegetationspunktes.
Berichte der Deutschen Botanischen Gesellschaft 32:77–
83. Lynott, RE, and OP Cramer. 1966. Detailed analysiss of the
1962
Columbus Day windstorm in Oregon and Washington.
Monthly Weather Review 94(2):105–117.
Lyons, LA. 1957. Insects affecting seed production in red
pine.
II. Dioryctria disclusa Heinrich, D. abietella (D. & S.) and
D. cambiicola (Dyar) (Lepidoptera: Phycitidae). Canadian
Entomologist 89(2):70–79.
Review of Forest Research Vol. 2, ed. JA Working Party on Douglas-fir Provenances S2.02–05,
Romberger and P Mikola. Academic Press, New Sept.3–5, Göttingen, Germany. HH Hattemer ed.
York. 315 p. Magnesen, S. 1978. Preliminary report on the international
Lyr, H, and V Garbe. 1995. Influence of root
temperature on growth of Pinus sylvestris, Fagus
sylvatica, Tilia cordata, and Quercus robus. Trees
9:220-223.
Lyr, H, H Polster, and HJ Fielder. 1967. Gehölzphysiologie.
Jena, VEB Gustav Fischer: 444.
MacDonald, J. 1952. The Place of North–Western
American Co- nifers in British Forestry. Sixth British
Commonwealth Forestry Conference Canada 1952.
Forestry Commission, London.
MacDonald, J. 1957. The place of exotic trees in British
silvi- culture, pp. 1–6 in Exotic Forest Trees in
Great Britain, ed. J Macdonald, RF Wood, MV
Edwards, and JR Aldhous. Forestry Commission
Bulletin No. 30, London, HMSO.
MacDonald, JE. 1996. Early development of bud
dormancy in conifer seedlings, pp. 193–199, in
Plant Dormancy: Physiology, Biochemistry and
Molecular Biology. ed. GA Lang. CAB
International, Wallingford, UK.
MacDonald, JE. 2000. The developmental basis of bud
dor- mancy in 1-year-old Picea and Pseudotsuga
seedlings, pp. 313–317 in Dormancy in Plants: From
Whole Plant Behaviour to Cellular Control, ed J-D
Viémont and J Crabbé. CAB International, Paris.
MacDonald, JE, and JN Owens. 1993a. Bud
development in coastal Douglas-fir seedlings
under controlled-envi- ronment conditions..
Canadian Journal of Forest Research 23:1203–
1212.
MacDonald, JE, and JN Owens, JN. 1993b. Bud
development in coastal Douglas-fir seedlings in
response to different dormancy-induction treatments.
Canadian Journal of Botany 71: 1280–1290.
MacDonald, JE, and JN Owens. 2006. Morphology,
physiology, survival, and field performance of
containerized coastal Douglas fir seedlings given
different dormancy-induction regimes.
HortScience 41(6):1416–1420.
MacDonald, JE, and JN Owens. 2010. Physiology and
growth of containerized coastal Douglas-fir
seedlings given different durations of short days to
induce dormancy. HortScience 45(3):342–346.
MacDougall, RS. 1906a. Megastigmus spermotrophus
Wachtl, as an enemy of Douglas fir (Pseudotstuga
taxifolia). Transac- tions of the Royal Scottish
Arboricultural Society 19:52–65, 2 pls.
MacDougall, RS. 1906b. A new enemy of the Douglas fir.
Journal of the Board of Agriculture [London] 12(10):615–621.
MacFall, JS. 1994. Effects of ectomycorrhizae on
biochemistry and soil structure, pp. 217–235 in
Mycorrhizae and Plant Health, ed. FL Pfleger and
RG Linderman, The American Phytopthological
Society, St. Paul, Minnesota. 344 pp.
Maciejowski, K. 1950. O pryzdatnosci daglezji dla
lasow pol- skich i o jej roli w gospodarstwie
lesnym (Douglas-fir in Polish forests, its
usefulness and importance to the forest
economy). Sylwan 94(1):58–75, (2):33–48.
Maciejowski, K. 1951. Egzoty naszych lasow (Exotic
trees in our woods). Panstwowe Wydawnictwo
Rolnicze i Lesne. Warsaw.
Macklin, C, and G Manning. 1992. British Columbia
Forestry Sector. Forestry Canada, Pacific Forestry
Centre, Victoria, B.C., Canada.
MacMorran, AM. 1946. Experimental Seed Storage. Royal
New Zealand Institute of Horticulture Journal 15 (3):
24-25. 1946.
Magnesen, S. 1973. Preliminary report on the
international short term Douglas-fir (Pseudotsuga
menziesii) provenance experiment in West Norway,
pp.126–139 in Proceedings Meeting of IUFRO
References 319
Ecological
short term Douglas-fir provenance experiment in West
Norway, pp. 257–259 in Vol.1: Background papers and
Douglas-fir Provenances. Proceedings IUFRO Joint Meet-
ing of Working Parties S2.02–05 Douglas-fir
Provenances, S2.02–06 Lodgepole Pine Provenances,
S2.02–12 Sitka spruce Provenances, S2.02–14 Abies
Provenances, Vancouver, BC. 1978. British Columbia
Ministry of Forests, Information Services Branch,
Victoria.
Magnesen, S. 1987. The international short term Douglas-
fir (Pseudotsuga menziesii) provenance experiment in
West Norway, pp. 25–30 in FBVA–Berichte No. 21.
Schriftenreihe der Forstlichen Bundesversuchsanstalt
Vienna, Austria.
Maguire, DA. 1994. Branch mortality and potential litter fall
from Douglas-fir trees in stands of varying density.
Forest Ecology and Management 70:41–53.
Maguire, DA, and JLF Batista. 1996. Sapwood taper models
and implied sapwood volume and foliage profiles for
coastal Douglas-fir, Canadian Journal of Forest
Research 26:849–862.
Maguire, DA, and WS Bennett. 1996. Patterns in vertical dis-
tribution of foliage in young coastal Douglas-fir.
Canadian Journal of Forest Research 26(11):1991–
2005.
Maguire, DA, SR Johnston, and J Cahilb. 1999. Predicting
branch diameter on second-growth Douglas-fir from
tree-level descriptors. Canadian Journal of Forest
Research 29:829–1940.
Maguire, DA, M Moeur, and WS Bennett. 1994. Models for
describing basal diameter and vertical distribution of
primary branches in young Douglas-fir. Forest Ecology
and Management 63:23–55.
Maguire DA, Kanaskie A, Voelker W, Johnson R, and G
Johnson. 2002. Growth of young Douglas-fir plantations
across a gradient in Swiss needle cast severity. Western
Journal of Applied Forestry 17(2):86–95.
Maguire, WP. 1955. Radiation, surface temperature, and
seedling survival. Forest Science 1:277–285.
Maguire, WP. 1956. Are ponderosa pines cone crops
predict-
able? Journal of Forestry, 54:778–779.
Mahadevan, A, N Raman, and K Natarajan. Mycorrhizae
For Green Asia. First Asian Conference on Mycorrhizae,
Madras, India 1988.
Maheshwari, P, H Singh. 1967. The female gametophyte of
gymnosperms. Biological Reviews 42(1):88–129.
Mair, AR. 1973. Dissemination of tree seed, Sitka spruce
Western hemlock and Douglas-fir. Scottish Forestry 27:
(4):308–314.
Majer, A. 1980. Vizsgálati eredmények a fafajmegválasztás
köreböl (Investigations on the choice of tree species).
EFE Közlemények 1:55–64.
Maki, DS, TM McDonough, and TL Noland, compliers. 1994.
Making the grade, an International Symposium on planting
stock performance and quality assessment, September 11–
15, 1994, Sault Ste. Marie, Ontario, Canada. IUFRO
S3.02-03, S1.05-04, S2.01-00. IUFRO and Ontario Forest
Research Institute, Sault Ste. Marie, Ontario. 95 p.
Malavasi, M de Matos, SG Stafford, and DP Lavender. 1985.
Stratifying, partially redrying and storing Douglas-fir
seeds: effects on growth and physiology during germina-
tion. Annals of Forest Science 42 (4) 371–384.
Malavasi, M de Matos, TM Ching, and DP Lavender.
1986. Stratifying, partially redrying, and storing
Douglas-fir seeds: biochemical responses. Annals of
Forest Science 43:35–48.
Mallett, KI. 1992. Armillaria root rot in the Canadian prai-
rie provinces. Information Report NOR-X-329,
Forestry Canada, Northwest Region, Northern
Forestry Centre, Edmonton, AB. 22 p.
Malloch, DW, KA Pirozynski, and PH Raven. 1980.
Marx, DH. 1972. Ectomycorrhizaes biological deterrents to
and evolutionary significance of mycorrhizal pathogenic root infections. Annual Reviews Phytopathol-
symbioses in vascular plants (A Review). ogy 10:429–454.
Proceedings of the National Academy of Sciences Marx, DH. 1973. Mycorrhizae and feeder root disease, pp.
USA 77(4):2113–2118. 351–382 in Ectomycorrhizae, Their Ecology and
Malyavkina, VS. 1958. (Spores and pollen of the Lower Physiology.
Cre- taceous of the eastern-Gobi depression). Trudy
VNIGRI No. 119. Leningrad. Partial reproduction
and transla- tion in Catalog of Fossil Spores and
Pollen, Vol. 28. Mezoic Megaspores, Microspores
and Pollen. A Travers, HT Ames, and GOW Kremp
(eds.) 1968. Palynology Laboratory, Pennsylvania
State University, State College Pennsylvania.
Manter, DK, PW Reeser, and JK Stone. 2005. A climate
based model for predicting geographic variation in
Swiss needle cast severity in the Oregon Coast
Range. Phytopathology 95:1256–1265.
Margolis, HA, and RH Waring. 1986. Carbon and
nitrogen allocation patterns of Douglas-fir seedlings
fertilized with nitrogen in autumn. II. Field
Performance. Canadian Journal of Forest Research
16:(5) 903–909.
Margules, SR. 1968. Douglas-fir, pp. 180–181 in
Growing Trees on Australian Farms.
Commonwealth of Australia, Dept. of National
Development, Forestry and Timber Bureau,
Canberra.
Margus, M. 1961. Ebatsuuga Eestis (Pseudotsuga in
Estonia). (With German summary “Die Douglasie
in Estland.”) Eesti NSV Teaduste Akadeemia juures
Asuva Loodusuurijate Aastaraamat 54:71–89.
Margus, M. 1963. Douglas-fir in Estonia. Lesnoe Khozjajstvo
9:24–27.
Marić, B. 1962. Izdvajanje semenskih sastojina
cetinara u nr srbiji (with French summary: A
propos de la sélection des peuplements à graine
des essences resineuses en Serbie). Topola
6(25/26):9–14.
Marković, L. 1950. Proucavanje razvoja vestacki
podignutih sastojina nekih cetinara na Avali (The
development of plantations of some conifer species
at Avala). Zbornik In- stituta za Naucna
Instrazuranja u Sumarstva Srbije 1:62–109.
Marosi, F. 1885. A külföldi fak meghonositásának
kérdéséhez (Remarks about the question of
acclimatization of exotic species). Erdészeti
Lapok 24:1171–1187.
Marshall, JD. 1986. Drought and shade interact to
cause fine root mortality in Douglas-fir seedlings.
Plant & Soil 91:51–60.
Marshall, KA, and DB Neale. 1992. The inheritance of
mito- chondrial DNA in Douglas-fir (Pseudotsuga
menziesii). Canadian Journal of Forest Research
22:73-75.
Martinez, M. 1949. Las Pseudotsugas de Mexico.
Anales del Instituto de Biologia 20(1/2):129–184.
Martinez, M. 1963. Las Pinaceas Mexicanas. 3rd ed. Universidad
Nacional Autonoma de Mexico, Mexico.
Martinsson, O. 1985. Douglas-fir Seed Collection for
Establishment of Progeny Trials in Sweden.
Institutionen för Skogsskötsel. Arbetsrapporter Nr.
2. Sveriges Lantbruksuniversitet, Umeå, Sweden.
Martinsson, O. 1990. Research on Douglas-fir in Sweden.
Sveriges Lantbruksuniversitet, Institutionen för
Skogsskötsel. Arbetsrapporter no. 42. Umeå,
Sweden.
Martinsson, O, and Kollenmark, R. 1993. Överlevnad
i pro- veniensförsök av Douglas (Pseudotsuga
menziesii (Mirb.) Franco) ett eller två år efter
plantering. (With English summary “Survival rate
in provenance trials of Douglas fir (Pseudotsuga
menziesii (Mirb.) Franco) one or two years after
planting”). Sveriges Landbruksuniversitet, Umeå,
Institutionen för Skogskdskötsel Arbetsrapporter
Nr.70.
320 Douglas-fir: The Genus Pseudotsuga
ed. GC Marks and TT Kozlowski. 444 pp. Academic pro- cess, pp. 51–57 in Plant Reproduction: From Floral
Press, New York. Induction to Pollination, ed. E Lord and G Bernier. American
Massicotte, H. 1994. Mycorrhizal fungi. Class Notes. Society
Silvicul- ture Institute of British Columbia, Module I:
Basic Principles. January 24–February 4, 1994, Green
Timbers, Prince George, BC.
Masters, CJ. 1982. Weyerhaeuser’s seed orchard program, pp
60–70 in Proc. 18th Can. Tree Improve. Assoc. Meeting,
Part
2. ed. DFW Pollard, DGW Edwards, and CW
Yeatman, Canadian Forest Service, Ottawa, ON.
Matthews, JD. 1953. Internal Report, Forestry Commission.
Origi- nal not seen, cited from R Lines, 1987, Forestry
Commis- sion Bulletin 66.
Matthews, JD. 1963. Factors affecting the production of seed
by forest trees. For. Abstr. 24(1):i-xiii.
Matthews, JD. 1983. The Role of Northwest American Trees
in Western Europe. HR MacMillan Lectureship in
Forestry, University of British Columbia, Vancouver,
BC.
Mattoon, WR. 1936. Forest Trees and Forest Regions of the
United States. USDA Forest Service Miscellaneous
Publication 217, Washington, DC.
Matson, WJ. 1978. The role of insects in the dynamics of
cone production of red pine. Oecologia (Berl.) 33:327–
349
Mátyás, C. 1992. Conservation problems of forest
ecosystems on the Great Hungarian Plain. Hungarian
Agricultural Review 1(1):33–37.
Mayr, H. 1907. Die Anbauversuche mit fremdländischen
Baumarten in den Staatswaldungen des Königreichs
Bay- ern. Forstwissenschaftliches Centralblatt 29:1–10,
129–137,
336–349.
McArdle, RE, and WH Meyer. 1930. The Yield of Douglas-
fir in the Pacific Northwest. USDA Forest Service
Technical Bulletin No. 201, Washington, DC. 64 p.
McArdle, RE, WH Meyer, and D Bruce. 1961. The yield of
Douglas-fir in the Pacific Northwest. UDSA Forest
Service, Technical Bulletin 201 (rev.). Washington, DC.
72 p.
McCaughey, WW, and T Weaver. 1991. Seedling
submergence tolerance of four western conifers. Tree
Planters' Notes 42(2):45–48.
McCloskey, SPJ, LD Daniels, JA McLean. 2009. Potential
Impacts of Climate Change on Western hemlock Looper.
Northwest Science 83(3):225–238.
McCorquodale, W. 1880. Abies Douglasii. Journal of Forestry
and Estate Management 4:362–364.
McCoy, EL, L Boersma, HL Ungs, and Akratanakul. 1984.
Toward Understanding soil water uptake by plant roots.
Soil Science 137(2):69–77.
McCreary, D, and ML Duryea. 1985. OSU vigor test, pp. 85–92
in Proceedings: Evaluating seedling quality: principles,
pro- cedures and predictive abilities of major tests.
Workshop held October 16–18, 1984. Forest Research
Laboratory, Oregon State University, Corvallis.
McCreary, DD, and ML Duryea. 1987. Predicting field
per- formance of Douglas-fir seedlings: Comparison
of root growth potential, vigor, and plant moisture
stress. New Forests 2:153–169.
McCreary, DD, Y Tanaka, and DP Lavender. 1978.
Regulation of Douglas-fir seedling growth and
hardiness by control- ling photoperiod. Forest Science
24(2):142–152.
McCreary, DD, DP Lavender, and RK Hermann. 1990. Pre-
dicted global warming and Douglas-fir chilling tempera-
tures. Annals of Forest Science 47:325–330.
McCulloch, WF. 1943. Ice breakage in partially cut and
uncut second growth Douglas-fir stands. Journal of
Forestry 41(4/ April):275–278.
McDaniel, CN. 1989. Floral initiation as a developmental
of Plant Physiologists, Rockville, Md. McKay, HM, and IMS White. 1997. Fine root electrolyte
McDaniel, CN, SR Singer, JS Gebhardt, and KA leak- age and moisture content: indices of Sitka spruce
Dennin. 1987. Floral determination: a critical and Douglas-fir seedling performance after desiccation.
process in meristem ontogeny, pp. 109–120, in The New Forests 13:139–162.
Manipulation of Flowering, ed. JG Atherton, McKay, HM, JR Aldhous, and WL Mason. 1994 Lifting,
Butterworth, London. stor- age, handling and dispatch, pp. 198–222 in Forest
McDermott, JM, and RA Robinson. 1989. Provenance Nursery
varia- tion for disease resistance in Pseudotsuga
menziesii to the Swiss needle-cast pathogen,
Phaeocryptopus gaeumannii. Canadian Journal of
Forest Research 19(2): 244–246.
McDonald, MB. 1993. The history of seed vigor testing.
Journal of Seed Technology 17(2):93–160.
McDonald, PM. 1990. Pseudotsuga macrocarpa (Vasey)
Mayr Big- cone Douglas-fir, pp. 520–526 in Silvics of
North America, Vol. 1. Conifers and Important
Hardwoods. USDA, Forest Service Agriculture
Handbook 654.
McDonald, PM, D Minore, and T Atzet. 1983.
Southwestern Oregon-northern California
hardwoods, pp. 29–32 in Silvicultural Systems for the
Major Forest Types of the United States. R Burns,
technical compiler. Agriculture Handbook 445,
USDA Forest Service, Washington, DC.
McDougall, WB. The classification of symbiotic phenomena.
Plant World 21:250–256.
McDowell, NG, J Licata, and BJ Bond. 2005.
Environmental sensitivity of gas exchange in
different sized trees. Oeco- logia 145:9–20.
McEvoy, T. 1943. Report of the first annual excursion to
the val- ley of the Suir, June 7–10th, 1943. Irish
Forestry 1(1):28–35.
McIntyre, GI. 2001. Control of plant development by
limiting factors: A nutritional perspective.
Physiologia Plantarum, 113(2):165–175.
McKay, HM. 1992. Electrolyte leakage from fine roots of
conifer seedlings: a rapid index of plant vitality
following cold storage. Canadian Journal of Forest
Research 22(9):1371–1377.
McKay, HM. 1993. Tolerance of conifer fine roots to cold
storage. Canadian Journal of Forest Research 23(3):337–342.
McKay, HM. 1994. Frost hardiness and cold storage
tolerance of the root system of Picea sitchensis,
Pseudotsuga menziesii, Larix kaempferi and Pinus
sylvestris bare-root seedlings. Scandinavian
Journal of Forest Research 9:203–213.
McKay, HM. 1997. A review of the effect of stresses
between lifting and planting on nursery stock
quality and perfor- mance. New Forests 13:369–
399.
McKay, HM. 1998. Root electrolyte leakage and root
growth potential as indicators of spruce and larch
establishment. Silva Fennica 32:241–252.
McKay, H, and B Howes. 1996. Recommended Plant
Type and Lifting Dates for Direct Planting and
Cold Storage of Bare- Root Douglas-fir in Britain.
Forestry Commission England, Research
Information Note 284, Forestry Commission,
Research Division, Farnham, Surrey, UK.
McKay, HM, and DC Malcolm. 1988. A comparison of
the fine root component of a pure and a mixed
coniferous stand. Canadian Journal of Forest Research
18(11):1416–1426.
McKay, HM, and WL Mason. 1991. Physiological
indicators of tolerance to cold storage in Sitka
spruce and Douglas-fir seedlings. Canadian Journal of
Forest Research 21(6):890–901.
McKay, HM, and AD Milner. 2000. Species and
seasonal vari- ability in the sensitivity of seedling
conifer roots to drying and rough handling.
Forestry 73(3):259–270.
McKay, HM, and JL Morgan. 2001. The Physiological
basis for the establishment of bare-root larch
seedlings. Forest Ecology and Management 142:1–
18.
References 321
April
Practice, ed. JR Aldhous and WL Mason. Forestry Com-
mission Bulletin 111, HMSO, London.
McKay, HM, BA Gardiner, WL Mason, DG Nelson, and MK
Hollingsworth. 1993. The gravitational forces generated
by dropping plants and the response of Sitka spruce
seedlings to dropping. Canadian Journal of Forest
Research 23:2443–2451.
McMichael, BJ, and JJ Burke. Soil temperature and root
growth.
HortScience 33(6):947–951.
McMinn, RG. 1963. Characteristics of Douglas-fir root
systems.
Canadian Journal of Botany 41:105–122.
McMullan, EE. 1980. Effect of applied growth regulators on
cone production in Douglas-fir, and relation of endog-
enous growth regulators to cone production capacity.
Canadian Journal of Forest Research 10(3):405–422.
McNabb, DH, and HA Froehlich. 1984. Conceptual model
for predicting forest productivity losses from soil
compaction, pp. 261–265 in New Forests for a Changing
World, Proceed- ings 1983 SAF National Convention,
Society of American Foresters, Portland, OR.
McNabb, DH, HA Froehlich, and F Gaweda. 1982. Aver-
age dry-season precipitation in Southwest Oregon, May
through September OSU Extension Service,
Miscellaneous Publication 8226. 7 p.
McWilliams, ERG, and G Thérien. 1997 (rev. ed.)
Fertilization and thinning effects on a Douglas-fir
ecosystem at Shawnigan Lake: 24-year growth response.
Canada-British Columbia Partnership Agreement on
Forest Resource Development: FRDA II Report No.
269, Canadian Forest Service, Pacific Forestry Centre,
British Columbia Ministry of Forests, Research Branch,
Victoria, BC. 44 p.
Means, JE, and ME Helm. 1985. Height growth and site
index curves for Douglas-fir on dry sites in the Willamette
National Forest. USDA Forest Service, Research Paper
PNW-341. Pacific Northwest Forest and Range
Experiment Station, Portland, OR. 17 p.
Meeks-Wagner, DR, ES Dennis, KT Thanh Van, and WJ
Pea- cock. 1989. Tobacco genes expressed during in
vitro floral initiation and their expression during
normal plant development. Plant Cell. 1(1):25–35.
Meilan R. 1997. Floral induction in woody angiosperms.
New Forests 14:179–202.
Meinzer, FC, BJ Bond, and JA Karanian. 2008. Biophysical
constraints on leaf expansion in a tall conifer. Tree
Physi- ology 28:197–208.
Mejnartowicz, L. 1973. Some results from the IUFRO
Douglas- fir experiment in Kornik near Poznan, pp.150–
157 in Proceedings Meeting of IUFRO Working Party on
Douglas-fir provenances S2.02–05, Sept.3–5. Göttingen,
Germany. HH Hattemer ed.
Mejnartowicz, L. 1976. Genetic investigation on Douglas-fir
(Pseudotsuga menziesii (Mirb.) Franco). Arboretum Kornickie
21:125–187.
Mejnartowicz, L, and A Lewandowski. 1994. Allozyme
poly- morphism in seeds collected from IUFRO–68
Douglas-fir test plantation. Silvae Genetica 43(4):181–
186.
Melchior, GH. 1960. Girdling experiments to increase
flower- ing ability in Japanese larch L. leptolepis (Sieb.
and Zucc.) Gord. and European larch L. deciduas Mill.
Silvae Genetica 9:106–111.
Mellerowicz, E, and M Bonnet-Masimbert. 1986.
Importance de la teneur en eau du pollen pour la
réalisation de croisements contrôlés chez le Douglas.
Annals of Forest Science 43, 179–188.
Mellerowicz, EJ, RT Riding, and MS Greenwood. 1995.
Nuclear and cytoplasmic changes associated with
maturation in the vascular cambium of Larix laricina.
Tree Physiology 15(7/8):443–449.
Menzies, A. 1923. Menzies’ Journal of Vancouver’s Voyage,
Management Paper 22, Simon Fraser University,
to October, 1792. Edited, with Botanical and Burnaby, BC. 138 p.
Ethnological Notes, by CF Newcombe, MD, and a Miller, GE. 1983. Biology, sampling and control of the
Biographical Note by J Forsyth. William H. Cullin, Douglas- fir cone gall midge, Contarinia oregonensis
Printer to the King’s Most Excellent Majesty, Foote (Diptera: Cecidomyiidae), in Douglas-fir seed
Victoria, BC. xx + 155 p. orchards in British Columbia. PhD dissertation, Simon
Merendi, A. 1956. La Douglasia. Italia Agricola Fraser University, Burnaby, BC. 192 p.
93(2):136–138. Merendi, A. 1965. Conviene coltivare Miller, GE. 1984a. Biological factors affecting Contarinia
orego- nensis (Diptera: Cecidomyiidae) infestations in
la Douglasia in Italia?
Douglas-fir
Italia Forestale e Montana 20(6):255–260.
Mergen, F. 1963. Ecotype variation Pine Ecology 44(1):716–727.
Merkle, R. 1951. Über die Douglasvorkommen und die
Ausb- reitung der Adelopus Nadelschütte in
Württemberg–Ho- henzollern. Allgemeine Forst–
und Jagdzeitung 122:161–192.
Merkle, SA, and WT Adams. 1987. Patterns of allozyme
varia- tion within and among Douglas-fir breeding
zones in Southwest Oregon. Canadian Journal of
Forest Research 17(5):402–407.
Merkle, SA, WT Adams, and RK Campbell. 1988.
Multivariate analysis of allozyme variation patterns
in coastal Douglas- fir from Southwest Oregon.
Canadian Journal of Forest Research 18(2):181–
187.
Meso, SW. 1979. Douglas-fir cone and seed insects, pp.
245–262 in Forest Insect Survey and Control, ed.
JA Rudinsky. Or- egon State University Book
Stores, Inc., Corvallis, OR.
Mexal, J, and Reid, CPP. 1973. The growth of selected
mycor- rhizal fungi in response to induced water
stress. Canadian Journal of Botany 51 (9):1579–
1588.
Meyer, BS, and DB Anderson. 1952. Plant Physiology. 2d ed.
D. Van Nostrand, New York.
Meyer, EA. 1914. Die Nadelhölzer im Arboretum des
land- wirtschaftlichen Institutes in Moskau.
Mitteilungen der Deutschen Dendrologischen
Gesellschaft 23:188–200.
Michaud, D. 1978. Résultats en pépiniere de 184
provenances
de douglas. Annales de recherches sylvicoles, de
l'AFOCEL
:41–87.
Michaud, D. 1987. First Results of American Douglas-fir
Provenance Trials in France. FBVA–Berichte No.
21:3–8. Schriftenreihe der Forstlichen
Bundesversuchsanstalt Vienna, Austria.
Michaud, D. 1997. Le matériel végétal, pp. 63–76 in Le Douglas.
AFOCEL, Paris.
Michaud, D, and A Najar. 1980. Étude du
débourrement de 181 provenances de Douglas.
Annales des Recherches Sylvicoles,
AFOCEL:153–193.
Miglioli, JA, and JJ Rozados. 1972. Informe sobre
crecimientos observados en coniferas exóticas en la
zona cordillerana del noroeste de la Provincia del
Chubut. With Engl. summary: The growth of exotic
conifers in the Cordil- lera region of the northwest
of the Province of Chubut, Argentina, pp. 1665–1669
in Proceedings 7th World Forestry Congress Buenos
Aires, Argentina, 4–18 Oct. 1972. Coité de
Documentación y Actas, Buenos Aires, 1978.
Miklavzic, J. 1951. Ozeleni duglazije. Gozdarski
covtnik, Ljubljana (Original not seen, cited from
Cokl 1965).
Miller, DJ. 1961. Permeability of Douglas-fir in Oregon.
Forest Products Journal 11(1):14–16.
Miller, DJ, and RD Graham. 1963. Treatability of
Douglas-fir from western United States. American
Wood–Preservers Association Proceedings 59:218–
222.
Miller, GE. 1980. Pest management in Douglas-fir
seed or- chards in British Columbia: a problem
analysis. Simon Fraser University Pest
322 Douglas-fir: The Genus Pseudotsuga
seed orchards on Vancouver Island, British Columbia. Bulletin No.
Environmental Entomology 13:873–877.
Miller, GE. 1984b. Pest management in Douglas-fir seed or-
chards in British Columbia, in Proceedings of the
IUFRO Cone and Seed Insects Working Party Conference.
Working Party S2.07–01. July 31–August 6, 1983,
Athens, GA, ed. HO Yates, III. USDA Forest Service
Southeastern Forest Experiment Station, Asheville, NC.
214 p.
Miller, GE. 1986a. Sampling major Douglas-fir cone and
seed insects in British Columbia, pp. 103–112 in
Proceedings of the 2nd Conference of the Cone and
Seed Insects Working Party S2.07-01 Briancon, 05100,
France, September 3-5, 1986, ed. A Roques. International
Union of Forestry Research Organizations, Olivet,
France.
Miller, GE. 1986b. Damage prediction for Contarinia orego-
nensis Foote (Diptera: Cecidomyiidae) in Douglas-fir
seed orchards. The Canadian Entomologist 118:1297–
1306.
Miller, GE. 1986c. Distribution of Contarinia oregonensis
Foote (Diptera: Cecidomyiidae) eggs in Douglas-fir seed
or- chards and a method of estimating egg density. The
Canadian Entomologist 118(12):1291–1295.
Miller, GE. 1986d. Insects and conifer seed production in the
Inland Mountain West: a review, pp. 225–237 in
Proceed- ings, Symposium on Conifer tree seed in the
inland moun- tain west Symposium, Missoula, August 5-
6, 1985. Ed. RC Shearer. USDA Forest Service General
Technical Report INT-203, Intermountain Forest and
Range Experiment Station, Ogden, UT.
Miller, GE, compiler. 1989. Proceedings of the 3rd Cone and
Seed Insects Working Party Conference, Working Party
S2.07-01, IUFRO, 26–30 June 1988, Victoria, BC. Forestry
Canada, Pacific Forestry Research Centre, Victoria, BC.
242 p.
Miller, GE, and AF Hedlin. 1984. Douglas-fir cone moth and
cone gall midge: Relation of damage and prolonged
diapause to seed cone abundance in British Columbia,
pp. 91–99 in Proceedings of the IUFRO Cone and Seed
Insects Working Party Conference. Working Party
S2.07–01. July 31–August 6, 1983, Athens, GA, ed. HO
Yates, III. USDA Forest Service Southeastern Forest
Experiment Station, Asheville, NC. 214 p.
Miller, GE, and DS Ruth. 1986. Effect of temperature during
May to August on termination of prolonged diapause in
the Douglas-fir cone moth (Lepidoptera: Tortricidae).
The Canadian Entomologist 118:1073–1074.
Miller, GE, and DS Ruth. 1989. The relative importance of
cone and seed insect species on commercially important
conifers in British Columbia, pp. 25–34 in Proceedings
of the 3rd Cone and Seed Insects Working Party
Conference, Working Party S2.07-01, IUFRO, 26–30 June
1988, Victoria, BC. GE Miller, comp. Forestry Canada,
Pacific Forestry Research Centre, Victoria, BC. 242 p.
Miller, GE, AF Hedlin, and DS Ruth. 1984. Damage by two
Douglas-fir cone and seed insects: correlation with cone
crop size. Journal of the Entomological Society of
British Columbia 81:46–60.
Miller, JM. 1914. Insect Damage to the Cones and Seeds of
Pacific Coast Conifers. Agricultural Bulletin 95, US
Department of Agriculture, Washington, DC. 7 p.
Miller, JM. 1915. Cone Beetles: Injury to Sugar Pine and
Western Yellow Pine. Agricultural Bulletin 243, US
Department of Agriculture, Washington, DC. 12 p.
Miller, JM. 1916. Oviposition of Megastigmus
spermotrophus in the seed of Douglas-fir. Journal of
Agricultural Research 6(2):65–68, illus.
Miller, JM, and JE Patterson. 1916. Manuscript report, June
1, 1916. (Original not seen, cited in Keen 1958).
Miller, JT, and FB Knowles. 1994. Introduced forest trees in
New Zealand. Recognition, role and seed source. 14.
Douglas-fir, Pseudotsuga menziesii (Mirbel) Franco. FRO
124. New Zealand Forest Research Institute Limited, Misra, S. 1994. Conifer zygotic embryogenesis, somatic em-
Rotorua, New Zealand. 38 pp. bryogenesis, seed germination; biochemical and molecu-
Miller, RE, and SR Webster, 1979. Fertilizer response in lar advances. Seed Science Research 4:357–384.
mature stands of Douglas-fir, pp. 126–132 in Mitchell, AF. 1983. Big trees in Scotland. Scottish Forestry
Proceedings, Forest Fertilization Conference, 1979
Sept. 25–27; Union, WA, ed. SP Gessel, RM
Kenady, and WA Atkinson. Contribution 40,
University of Washington College of Forest
Resources, Institute of Forest Resources, Seattle.
Miller, RE, HW Anderson, and DC Young. 1988. Urea
and Biuret stimulate growth of Douglas-fir and
western hemlock seedlings. Soil Science Society of
America Journal 52(1):256–260.
Miller, RE, PR Barker, CE Peterson, and SR Webster.
1986. Using nitrogen fertilizer in management of
coast Doug- las-fir: I. Regional trends and response,
pp. 290–303 in Douglas-fir: Stand management for
the future, ed. CD Oliver, DP Hanley, and JA
Johnson. Contribution 55, University of
Washington, College of Forest Resources, Institute
of Forest Resources, Seattle.
Milliron, HE. 1949. Taxonomic and biological
investigations in the genus Megastigmus with
particular reference to the taxonomy of the Nearctic
species (Hymenoptera: Chalcidoidea:
Callimomidae). American Midland Natural- ist 41:
257–420.
Minelli, H. 1967. Bericht über die in Niederösterreich
vorkom- menden fremdländischen Baumarten.
Centralblatt für das gesamte Forstwesen 84(1):35–
63.
Ministère de l’Agriculture. 1958. 1950, Forêts, Chasse,
Pêche. Premier rapport général de l’Administration
des Eaux et Forêts, Brussels.
Minnich, RA. 1982. Pseudotsuga macrocarpa in Baja California?
Madroño 29(1):22–31.
Minore, D. 1970. Seedling Growth of Eight
Northwestern Tree Species Over Three Water
Tables. USDA Forest Service Pacific Northwest
Forest and Range Experiment Station Research
Note PNW-115 8 p.
Minore, D, and CE Smith. 1971. Occurrence and
growth of four northwestern tree species over
shallow water tables. USDA Forest Service,
Research Note PNW-160. Pacific Northwest Forest
and Range Experiment Station, Portland, OR. 9 p.
Minore, D. 1972. Germination and Early Growth of
Coastal Tree Species of Organic Seed Beds. USDA
Forest Service Research Paper PNW-134. Pacific
Northwest Forest and Range Experiment Station,
Portland, Oregon 20 p.
Minore, D. 1979. Comparative autecological
characteristics of northwestern tree species: a
literature review. USDA For- est Service, General
Technical Report PNW-87. Pacific Northwest
Forest and Range Experiment Station, Port- land,
OR. 72 p.
Minore, D, and D Kingsley. 1983. Mixed conifers of
south- western Oregon, pp. 23–25 in Silvicultural
systems for the major forest types of the United
States, RM Burns, technical coordinator.
Agricultural Handbook 445, USDA Forest Service,
Washington, DC.
Minore, D. 1988. Effects of light intensity and
temperature on the growth of Douglas-fir and
incense-cedar seedlings. Forest Science 34:215–
223.
Mirov, NT. 1956. Photoperiod and flowering of pines.
Forest Science 2:32–332.
Misra, S. 1993. SDS-PAGE and Western blot analysis of
storage protein motilization following germination of
Douglas-fir seeds, pp. 73–78 in Dormancy and
barriers to germination: Proceedings of an
International Symposium of IUFRO Project Group P2
04 -00, Seed Problems, ed. DGW Edwards. Forests,
Canada, Victoria, BC. 153 p.
References 323
glasia (Pseudotsuga taxifolia (Poir.)Britt.), pp. 51–73 in
37(4):288–292.
Mitchell, RG. 1974. Estimation of needle populations on
young, open-grown Douglas-fir by regression and life
table analysis. USDA Forest Service Research Paper
PNW-181, Pacific Northwest Forest and Range
Experiment Station, Port- land, OR. 14 p.
Modert, P. 1965. Les régions forestières du Grand-Duché de
Luxembourg. Bulletin de la Société Royale de Botanique
de Belgique 72(3):125–166, (4):187–213.
Molina, R. 1981. Mycorrhizal inoculation and its potential
impact on seedling survival and growth in southwest Ore-
gon, pp. 86-91 in Reforestation of Skeletal Soils:
Proceedings of a Workshop, ed. SD Hobbs and OT
Helgerson. Forest Research Laboratory, Oregon State
University, Corvallis.
Molina R, JE Smith, D McKay, and LH Melville. 1997.
Biology of the ectomycorrhizal genus, Rhizopogon III.
Influence of co-cultured conifer species on mycorrhizal
specificity with the arbutoid hosts Arctostaphylos uva-
ursi and Arbutus menziesii. New Phytologist 137:519–
528.
Molina R, JM Trappe, LC Grubisha, and JW Spatafora. 1999.
Rhizopogon, pp. 129–161 in Ectomycorrhizal Fungi:
Key Genera in Profile, ed. JWG Cairney and SM
Chambers. Springer-Verlag, Heidelberg.
Moller, C, E Falleri, E Laroppo, and M Bonnet-Masimert.
1999. Drying and storage of prechilled Douglas-fir
(Pseu- dotsuga menziesii) seeds. Canadian Journal of Forest
Research 29:172–177.
Moltesen, P. 1988. Skovtraernes ved og deres anvendelse.
Skovteknisk Institut, Copenhagen.
Monleon, VJ, M Newton, C Hooper, and JC Tappeiner II.
1999. Ten year growth response of young Douglas-fir to
vari- able density Varnish leaf Ceanothus and herb
competition. Western Journal of Applied Forestry
14(4):208–212.
Monserud, RA. 1984. Height growth and site index curves
for inland Douglas-fir based on stem analysis data and
forest habitat type. Forest Science 30(4):943–965.
Monserud, RA, and GE Rehfeldt. 1990. Genetic and environ-
mental components of variation of site index in inland
Douglas-fir. Forest Science 36(1):1–9.
Montville, ME, and DL Wenny. 1990. Application of foliar
fertil- izer during bud initiation treatments to container-
grown conifer seedlings. Rocky Mountain Experiment
Station General Technical Report TRM 200, USDA Forest
Service, Fort Collins, CO, pp. 233–239.
Moore, AW. 1940. Wild Animal Damage to Seed and Seedlings
on Cut-over Douglas-fir Lands of Oregon and Washington.
Tech- nical Bulletin 706, USDA Forest Service,
Washington, DC.
Moore, JA. 1988. Response of Douglas-fir, grand fir, and
western white pine to forest fertilization, pp. 226–230
in Proceedings-future forests of the Mountain West: A
stand culture symposium. Missoula, MT, Sept. 29–Oct.
3, 1986, CW Schmidt, comp. USDA Forest Service,
General Tech- nical Report INT-243, Intermountain
Research Station, Ogden, UT.
Moore, J, and B Gardiner. 2001. Relative windfirmness of
New Zealand-grown Pinus radiata and Douglas-fir: A
preliminary investigation. New Zealand Journal of
Forest Science 31(2):208–223.
Moore, TC. 1979. Biochemistry and Physiology of Plant
Hormones.
Springer-Verlag, New York.
Moran, GF, and WT Adams. 1989. Microgeographical
patterns of allozyme differentiation in Douglas-fir from
Southwest Oregon. Forest Science 35(1):3–15.
Morandini, R. 1961. Indirizzi di azione technico–economici per
l’espansione delle specie forestali e della loro utilizzazione
industriale. Camera di Commercio Industria e Agricoltura
Trento «Economia Trentina» No. 1–2. Reprint
Morandini, R. 1968. Esperienze sulle provenienze della Dou-
Trees of the United States. USDA Miscellaneous
Studie e Ricerche di Genetica Forestale. Publication 287, USDA Forest Service, Washington, DC.
Pubblicazioni dell’ Istituto Sperimentale per la Muren, RC, TM Ching, and KK Ching. 1979. Metabolic
Selvicoltura Arezzo, No. 15. study of Douglas-fir pollen germination in vitro.
Morandini, R. 1968. Selezione individuale della Physiologia Plantarum 46: 287–292.
Douglasia (Pseudotsuga taxifoliaa Poir. Britt.), pp.
17–50 in Studi e Ricerche di Genetica Forestale.
Pubblicazioni dell’ Instituto Sperimentale per la
Selvicoltura Arezzo, No. 15.
Mörmann, P. 1956a. Zum Doulgasienanbau in
Nordbaden. Feststellungen und Erkenntnisse
anlasslich der Doug- lasienzapfenernte 1953.
Allgemeine Forst und Jagdzeitung 127(5/6):97–102.
Mörmann, n.i. 1956b. Erfahrungen in der Exotenanzucht.
Allgemeine Forstzeitschrift 11(8/9):116–118.
Morris JW, P Doumas, RO Morris, and JB Zaerr. 1990.
Cytoki- nins in vegetative and reproductive buds of
Pseudotsuga menziesii. Plant Physiology 93(1):67–
71
Morris, WG. 1936. Viability of conifer seed as affected
by seed- moisture content and kiln temperature.
Journal of Agricultural Research 52:855–864
Morris, WR, RR Silen, and H. Irgens-Møller. 1957.
Consis- tency of bud bursting in Douglas-fir.
Journal of Forestry 55:208–210.
Morrison, PH, and FJ Swanson. 1990. Fire history and
pattern in a Cascade Range landscape. Pacific
Northwest Research Station General Technical
Report PNW-GTR-254, USDA Forest Service. 77
p.
Moser, JC, RL Smiley, and IS Otvos. 1987. A new
Pyemotes (Acari: Pyemotidae) reared from the
Douglas-fir cone moth. International Journal of
Acarology 13(2):141–147.
Moss, EH. 1944. The prairie and associated vegetation of
Southwestern Alberta. Canada Journal Research 22:11–31.
Mousseaux, MR, RK Dumbroese, RL James, DL
Wenny, and GR Knudsen. 1998. Efficiency of
Trichoderma harzianum as a biological control of
Fusarium oxysporum in container- grown Douglas-
fir seedlings. New Forests 15:11–21.
Muhs, H-J. 1974. Distinction of Douglas-fir
provenances us- ing peroxidase–isoenzyme–
patterns of needles. Silvae Genetica 23(1/3):71–
76.
Muldavin, EH, RL DeVelice, and F Ronco, Jr. 1996. A
classifica- tion of forest habitat types southern
Arizona and portions of the Colorado Plateau. USDA
Forest Service, General Technical Report RM-GTR-
287, Rocky Mountain Research Station, Ogden, UT.
Muller C, E Falleri, E Laroppe, and M Bonnet-
Masimbert. 1999. Drying and storage of prechilled.
Douglas-fir, Pseudotsuga menziesii, seeds.
Canadian Journal of Forest Resources 29:172–177.
Münch, E. 1923. Anbauversuche mit Douglasfichten
verschie- dener Herkunft and anderen
Nadelholzarten. Mitteilungen der Deutschen
Dendrologischen Gesellschaft 33:61–79.
Münch, E. 1924. Nachtrag zu meinem Anbauversuch
mit Douglasfichten verschiedener Herkunft.
Mitteilungen der Deutschen Dendrologischen
Gesellschaft 34:373.
Münch, E. 1928. Klimarassen der Douglasie.
Centralblatt für das gesamte Forstwesen
54(1):254–260.
Münch, E. 1932. Die Douglasienschütte im Rheinland.
Der Deutsche Forstwirt 14(64):443–444.
Munger, TT. 1946. Windfall in relation to cutting.
Pacific Northwest Forest and Range Experiment
Station Research Note No. 34, USDA Forest
Service, Portland, OR. 12 pp.
Munger, TT, and WG Morris. 1936. Growth of
Douglas-fir Trees of Known Seed Source. Technical
Bulletin 537, USDA Forest Service, Washington,
DC.
Munns, EN. 1938. The Distribution of Important Forest
324 Douglas-fir: The Genus Pseudotsuga
Murray, A. 1884. Abies Douglasii, pp. 115–134 in Vol. 2 of Portland, OR. 5 p.
The Pinetum Britanicum: a Descriptive Account of Hardy Nelson, RE. 1965. A Record of Forest Plantings in Hawaii. USDA,
Conifer- ous Trees Cultivated in Great Britain. W. Forest Service, Resource Bulletin PSW–1.
Blackwood and Sons and Edward Ravenscroft,
Edinburgh and London (irreg. pag.).
Nakamura, RR, and NC Wheeler. 1992. Pollen competition
and
paternal success in Douglas-fir. Evolution 46(3):846-
851.
Namkoong, G, and HH Kang. 1990. Quantitative genetics
of
forest trees. Plant Breeding Review 8:1.39–188.
Nambiar, EKS. 1981. Ecological and physiological aspects of
the development development of roots: From nursery
to forest, pp. 117–29 in Proceedings of the Australian
For- est Nutrition Workshop, Productivity in Perpetuity,
CSIRO, Melbourne, 366 p.
Nanson, A. 1964. Enquête sur la résistance de diverses prove-
nances de Douglas vert à l'hiver 1962–1963 en Belgique.
Bulletin de la Société royale forestière Belgique 71:1–11.
Nanson, A. 1973. International Douglas-fir provenance experi-
ment in Belgium. Early results, pp. 158–172 in
Proceedings Meeting of IUFRO Working Party on
Douglas-fir Provenances S2.02–05, Sept.3–5, Göttingen,
Germany, ed. HH Hattemer.
Nanson, A. 1978. Belgian provenance experiments with Doug-
las-fir, grand fir, and Sitka spruce, pp. 335–346 in Vol. 1:
Background and Douglas-fir provenances. Proceedings of
the IUFRO Joint Meeting of Working Parties S2.02–05
Douglas-fir Provenances, S2.02–06 Lodgepole Pine
Provenances, S2.02–12 Sitka Spruce Provenances, S2.02–14
Abies Provenances, Van- couver Canada 1978. British
Columbia Ministry of Forests, Information Services
Branch, Victoria, BC.
Navasajtis, MZ. 1966. (Biology of flowering and seed–bear-
ing of Pseudotsuga taxifolia var. glauca and Picea
glauca in Lithuania). Lietuvos TSR Mokslu Akademijos
Darbai Serija C. 2(40):3–20.
Neale, DB, NC Wheeler, and RW Allard. 1986. Paternal in-
heritance of chloroplast DNA in Douglas-fir. Canadian
Journal of Forest Research 16:1152–1154.
Nebeker, TE. 1977. A partial life table for the Douglas-fir cone
moth, Barbara colfaxiana (Lepidoptera: Olethreutidae).
The Canadian Entomologist 109(7):943–951.
Nedkvitne, K. 1964. En vurdering av dyrkningsverdien til
utenlandske bartrearter på Vestlandet og deira plass i
landsdelen sitt skogbruk i framtida (An estimate of the
value of some exotic conifers for cultivation in West
Nor- way and their place in the future forestry of the
region). Norsk Skogbruk 10(13/14):385–390.
Neger, F. 1914. Der Stand der Anbauversuche mit
fremdlän- dischen Holzarten in den Staatswaldungen
des Köni- greiches Sachsen. Naturwissenschaftliche
Zeitschrift für Forst– und Landwirtschaft 12(1):1–9.
Nelson, EA, and DP Lavender. 1979. The chilling
requirement of Western hemlock seedlings. Forest
Science 25:485–490.
Nelson, EE, and RN Sturrock. 1993. Susceptibility of
Western Conifers to laminated root rot (Phellinus weirii)
in Oregon and British Columbia field tests. Western
Journal of Applied Forestry 8(2):67–70.
Nelson, EE, NE Martin, and RE Williams. 1981. USDA
Forest Service Forest Insect and Disease Leaflet 159, 6
p.
Nelson, EE, MG McWilliams, and WG Thies. 1994.
Mortality and growth of area-fertilized Douglas-fir on a
Phellinus- weirii infested site in Oregon. Western
Journal of Applied Forestry 9(2):52–56.
Nelson, EE, WG Thies, and CY Li. 1986. Are Seed and Cone
Pathogens Causing Significant Losses in Pacific Northwest
Seed Orchards? USDA Forest Service Pacific Northwest
Research Station. Research Note PNW-RN-436,
Nelson, S. 1976. Douglas-fir grove grows in Hawaii. Nooden, LD, and JA Weber. 1978. Environmental and hor-
The Sunday Oregonian, September 5, 1976.
Neustein, SA. 1970. Natural Regeneration of Conifers in
the UK and Eine. Conifer Conference 1970.
Neumann, A. 1993. A National Exotic Forest
Description as at 1 April 1991. Ministry of Forest
Report, Wellington, New Zealand.
New Zealand Forest Research Institute Rotorua. 1981.
Douglas- fir Seed Sources. Report 1–1 to 12–31,
1980:11–12. New Zealand Forest Service,
Wellington.
New Zealand. Annual Report of the Director of
Forestry for the year ended 31st March 1943.
Government Printer, Wellington. 1943. 24 p.
Nicholson, JJM. 1963. A Provenance Study of Douglas-fir
Grown under Controlled Light and Temperature
Conditions. BSc the- sis, Faculty of Forestry,
University of British Columbia, Vancouver.
Nielsen, KF, and EC Humphries. 1966. Effects of root
tem- perature on plant growth. Soils Fertilization
29:1–7.
Nightingale, GT. 1935. Effects of temperature on
growth anatomy and metabolism of apple and peach
roots. Bo- tanical Gazette 96(4):581–639.
Niklas, KJ. 1984. The motion of windborne pollen
grains around ovulate cones: implication on wind
pollination. American Journal of Botany 71: 356–
374.
Niklas, KJ. 1985a. The aerodynamics of wind pollination.
Botanical Review 51: 328–386.
Niklas, KJ. 1985b. Wind pollination—a study in controlled
chaos. American Scientist 73: 462–470.
Niklas, KJ. 1987. Aerodynamics of wind pollination. Scientific
American July: 90–95.
Niklas, KJ. 1992. Plant biomechanics: an engineering
approach to plant form and function. University of
Chicago Press, Chicago.
Niklas, KJ, and KT Paw U. 1982. Pollination and
airflow pat- terns around conifer ovulate cones.
Science 217: 442–444.
Niklas, KJ, and KT Paw U. 1983. Conifer ovulate cone
morphol- ogy: implications on pollen impaction
patterns. American Journal of Botany 70: 568–577.
Nikojaeva, MG. 1977. Factors controlling the seed
dormancy pattern, pp. 51–74 in The Physiology and
Biochemistry of Seed Dormancy and Germination.
ed. AA Khan, Newtock, North Holland,
Amsterdam.
Nikolaeva, MB. 1969. Physiology of Deep Dormancy in
Seeds. 12 datel’s ve “Nacha”, Leningrad. Translated
from Russian by Z. Shapiro, National Science
Foundation Washington, DC.
Nitsch, JP. 1957. Growth responses of woody plants to
photo- periodic stimuli. Proceedings of the American
Horticultural Society 70:512–525.
Nitsch, JP. 1965. Physiology of flower and fruit
development, pp. 1537–1647 in Encyclopedia of
Plant Physiology, ed. W. Ruhland. Springer-Verlag,
Berlin.
Niwa, CG, and DL Overhulser. 1992. Oviposition and
devel- opment of Megastigmus spermotrophus
(Hymenoptera: Torymidae) in unfertilized Douglas-
fir seed. Journal of Economic Entomology
85(6):2323–2328.
Nobbe, F. 1895. Welche fremdländischen Holzarten
können für Sachsen Bedeutung gewinnen.
Sächsischer Forstver- einsbericht 1895:1–17.
Noel, ARA. 1970. The girdled tree. Botannical Reviews
36:162– 193.
Noland, TL, GH Mohammed, and M Scott. 1997. The
depen- dence of root growth potential on light
level, photosyn- thetic rate, and root starch
content in jack pine seedlings. New Forests
13:105–119.
References 325
before storage: testing the model, pp. 195–200 in Target
monal control of dormancy in terminal buds of plants, Seedling Symposium, ed. R Rose, SJ Campbell, and TD Lan-
pp. 221–268 in Dormancy and Developmental Arrest. dis. USDA Forest Service General Technical Report. 286 p.
ed. ME Clutter, Academic Press, New York. O’Neill, GA. 1999. Genetics of fall, winter, and spring cold
North, AW. 1907. The uncharted Sierra of the San Pedro
Martir.
Bulletin of the American Geographical Society 39:544–
554. Northwest Forest Tree Seed Committee. 1959. Rules for
Service
Testing Forest Tree Seed of the Pacific Northwest.
Agriculture
Experiment Station, Oregon State College, Corvallis,
OR. 27 p.
Norum, RA. 1975. Characteristics and effects of
understory fires in western larch/Douglas-fir stands.
Doctoral dis- sertation, University of Montana,
Missoula. 155 p.
Nozicka, J. 1963. Zavádeni douglasky v ceskych zemich do
R 1918 (with English summary: Introduction of
Douglas- fir in Bohemia and Moravia). Práce VULHM
27:207–242.
O’Brien, IEW, DR Smith, RC Gardner, and BG Murray.
1996. Flow cytometric determination of genome size in
Pinus. Plant Science 115:91–99.
Odén, PC, Q Wang, KA Högberg, and M Werner. 1995.
Trans- port and metabolism of gibberellins in relation
to flower bud differentiation in Norway spruce (Picea
abies). Tree Physiology 15:451–456.
O’Driscoll, J. 1973. IUFRO Douglas-fir provenance experiment
in Ireland, pp.173–187 in Proceedings Meeting of IUFRO
Working Party on Douglas-fir Provenances S2.02–05, Sept.3–
5, Göttingen, Germany, ed. HH Hattemer.
O’Driscoll, J. 1978. Six year phenological study of thirty–two
IUFRO provenances of Douglas-fir, pp. 265–295 in Vol. 1:
Background papers and Douglas-fir provenances,
Proceedings IUFRO Joint Meeting of Working Parties S2–02–
05 Douglas-fir Provenances, S2–02–06 Lodgepole Pine
Provenances, S2–02–12 Sitka Spruce Provenances, S2–02–
14 Abies Provenances, Vancouver, Canada 1978. British
Columbia Ministry of Forestry, Information Services
Branch, Victoria, BC.
Oeschger, HJ. 1973. Gefährdung der Douglasie durch Frost-
trocknis? Allgemeine Forstzeitschrift 28(9/10):190-191.
Oeschger, HJ. 1975. Douglasienanbau in Baden–
Württemberg unter besonderer Berücksichtigung der
geschichtlichen Ent- wicklung. Schriftenreihe der
Landesforstverwaltung Baden–Württemberg. Vol. 45.
O’Flanagan, LP. 1968. Inventory of Woodlands of the Forest
and Wildlife Service. Forest and Wildlife Service,
Dublin, Ireland (Original not seen, cited from O’Driscoll
1978).
Ohwi, J, FG Meyer, and EH Walker. 1965. Flora of Japan.
Smithsonian Institute, Washington, D.C.
Oksbjerg, E. 1965. Three Conifer Species in Central Jutland.
Hedeselsk. Tidsskrift 86(6):113–128. (Original not seen,
cited from For. Abst. 27(2):2618, 1966).
Okubo H. 2000. Growth cycle and dormancy in plants, pp. 1–
22 in Dormancy in Plants: From Whole Plant Behaviour
to Cellular Control, ed J-D Viémont and J Crabbé. CAB
International, Paris.
Oliveira, JAL. 1967. Essências exóticas na rearborização das
serras do Norte de Portugal. Direcção–Geral dos Servicos
Florestais e Aquícolas, Lisbon, Portugal. Estudos e
Informação No. 232.
Olson, DL, and RR Silen. 1975. Influence of Date of Cone
Collec- tion on Douglas-fir Seed Processing and
Germination. USDA Forest Service Research Paper.
PNW 190. 11 p.
Olson, JS, and H Nienstaedt. 1959. Eastern Hemlock Seeds
and Seedlings, Response to Photoperiod and Temperature.
Con- necticut Agricultural Experiment Station Bulletin
620, New Haven.
Omi, S, and R Rose. 1990. Target root starch concentrations
under natural and cone-inducing conditions. Forest
hardiness in coastal Douglas-fir seedlings. Ecology and Management 19(1-4):85–97.
Doctoral dis- sertation, Oregon State University, Owens, JN. 1991. Flowering and seed set, pp. 247–271 in
Corvallis 80 p. Physiology of Trees, ed. AS Raghavendra. John Wiley
O’Neill, GA, WT Adams, and SM Aitken. 2001. and Sons Inc., New York.
Quantitative genetics of spring and fall cold
hardiness in seedlings from two Oregon
populations of coastal Douglas-fir. Forest
Ecology and Management 149:305–318.
O’Neill, GA, SW Aitken, and WT Adams. 2000.
Genetic selec- tion for cold hardiness in coastal
Douglas-fir seedlings and saplings. Canadian Journal
of Forest Research 30:1799–1807.
Oppermann, A. 1922. Den grønne douglasies vaekst i
Danmark (with Engl summary: The Douglas fir in
Denmark). Det forstlige Forsøgsvaesen i Danmark
6:350–360.
Oppermann, A. 1929. Racer af Douglasie og Sitkagran.
Det forstlige Forsøgsvaesen i Danmark 10:85–178.
O’Reilly, C, N McCarthy, M Keane, CP Harper, and
JJ Gar- diner. 1999. The physiological status of
Douglas-fir seed- lings and the field performance
of freshly lifted and cold stored stock. Annals of
Forest Science 56:297–306.
Orr-Ewing, AL. 1965. Inbreeding and single crossing in
Doug- las-fir. Forest Science 11(3):279–290.
Orr-Ewing, AL. 1966a. Frost damage to immature
Douglas- fir cones. Forest Research Review,
British Columbia Forest Service, pp. 77–80.
Orr-Ewing, AL. 1966b. Inter- and intraspecific crossing
of Douglas-fir (Pseudotsuga menziesii (Mirb.)
Franco). Silvae Genetica 15(4):121–126.
Orr-Ewing, AL. 1973. The Douglas-fir arboretum at
Cowichan Lake, Vancouver Island. British Columbia
Forest Service Research Note 57. 47p.
Orr-Ewing, AL, Frazer, AR, Karlsson, I. 1972. Interracial
crosses with Douglas-fir, early field results. BC Forest
Service Re- search Note 55. 33 p.
Osborne, DJ. 1983. Biochemical control systems
operating in the early hours of germination.
Canadian Journal of Botany 61:3568–2577.
Oswald, DD, F Hegyi, and A Becker. 1986. The current
status of coast Douglas-fir timber resources in
Oregon, Washing- ton, and British Columbia, pp.
26–31 in Douglas-fir: Stand Management for the
Future, ed. CD Oliver, DO Hanley, and JA
Johnson. Institute of Forest Resources, Contribu-
tion No. 55, College of Forest Resources,
University of Washington, Seattle.
Oswald, H, and J Pardé. 1984. Die Douglasie in Frank-
reich: Waldbau und Leistung. Allgemeine
Forstzeitung 39(17/18):438–441.
Overton, WS, and KK Ching. 1978. Analysis of
differences in height growth among populations in a
nursery selection study of Douglas-fir. Forest
Science 24(4):497–509
Overton, WS, DP Lavender, and RK Herman. 1973.
Estimation of Biomass and Nutrient Capital in
Stands of Old-Growth Douglas-Fir. IUFRO Biomass
Studies, S4.01, S4.01 Mensu- ration, Growth and
Yield, International Union of Forest Research
Organizations. University of Maine, Orono, ME,
pp. 89–103.
Owens, JN. 1968. Initiation and development of
leaves in
Douglas-fir. Canadian Journal of Botany 46:271–278.
Owens, JN. 1984a. Bud development in mountain hem-
lock (Tsuga mertensiana). Cone-bud differentiation
and predormancy development. Canadian Journal
of Botany 62:484–494.
Owens, JN. 1984b. Bud development in grand fir (Abies
gran- dis). Canadian Journal of Forest Research
14:575–588.
Owens, JN. 1987. Development of Douglas-fir apices
326 Douglas-fir: The Genus Pseudotsuga
Owens, JN, and MD Blake. 1985. Forest Tree Seed Production. Pankow, W, Boller, T, and Wiemken. 1991. Structure,
Canadian Forestry Service, Information Report PI-X- function, and ecology of the mycorrhizal symbiosis:
53. Petawawa National Forestry Institute, Chalk River, Introduction. Experientia 47:311–312.
Ontario. 161 p. Pankow, W, Boller, T, and Wiemken. 1991. The significance
Owens, JN, and M Molder. 1971. Pollen development in of mycorrhizas for protective ecosystems. Experientia
Douglas-fir (Pseudotsuga menziesii). Canadian Journal 47:391–394, Birkhauser Verlag CH-4010
of Botany 49:1263–1266. Basel/Switzerland.
Owens, JN, and M Molder M. 1973. A study of DNA and Papp, L. 1961. A duglásfenyö jelentösége
mitotic activity in the vegetative apex of Douglas-fir Magyarországon (The importance of Douglas-fir in
dur- ing the annual growth cycle. Canadian Journal of Hungary). Erdö 10(12):529–537.
Botany 51:1395–1409. Pardé, J. 1956. Douglas et tables de production. Annales de
Owens, JN, and SJ Morris. 1988. An ultrastructural study of l’Ecole Nationale des Eaux et Forêts 15(1):137–168.
fertilization in Douglas fir [Pseudotsuga menziesii Pardé, J. 1985. Die Douglasie in Ostfrankreich. Allgemeine
(Mirb.) Franco], pp. 339–344 in Sexual Reproduction in Forstzeitung 40(4):62–64, 80.
Higher Plants, Proceedings of the Tenth International
Symposium on the Sexual Reproduction in Higher Plants, Parke, JL. 1985. Effects of environment on mycorrhizae:
30 May–4 June 1988 University of Siena, Siena, Italy, ed. leav- ing the Dark Ages, pp. 107–109 in Proceeding 6
M Cresti, P Gori, and E Pacini. Springer-Verlag, Berlin NACH, June 25–29, 1984, Bend, OR, ed. R Molina. OSU
Heidelberg. Forest Research Laboratory, College of Forestry, Oregon
State University, Corvallis.
Owens, JN, and SJ Morris. 1990. Cytological basis for cyto-
plasmic inheritance in Pseudotsuga menziesii. I. Pollen Parke, JL, RG Linderman, and JM Trappe. 1983a. Inocu-
tube and archegonial development. American Journal of lum potential of ectomycorrhizal fungi in forest soils
Botany 77:433–445. of southwest Oregon and northern California. Forest
Science 30(2) 300–304.
Owens, JN, and SJ Morris. 1991. Cytological basis for
cytoplas- mic inheritance in Pseudotsuga menziesii. 11. Parke, JL, RG Linderman, and CH Black. 1983b. The role of
Fertilization and proembryo development. American ectomycorrhizas in drought tolerance of Douglas-fir
Journal of Botany 78:1515–1527. seedlings. New Phytologist 95:83–95.
Owens, JN, and SJ Simpson. 1982. Further observations on Parke, JL, RG Linderman, and JM Trappe. 1983c. Effects of
the pollination mechanism and seed production of forest litter on mycorrhiza development and growth of
Douglas- fir. Canadian Journal of Forest Research Douglas-fir and western red cedar seedlings. Canadian
12:431–434. Journal of Forest Research 13(4):666–671.
Owens, JN, Ø Johnsen, OG Dæhlen, and T Skrøppa. 2001. Parke, JL, RG Linderman, and JM Trappe. 1983d. Effect of
Potential effects of temperature on early reproductive root zone temperature on ectomycorrhiza and vesicu-
development and progeny performance in Picea abies lar–arbuscular mycorrhiza formation in disturbed and
(L.) Karst. Scandinavian Journal of Forest Research undisturbed forest soils of southwest Oregon. Canadian
16(3):2001. Journal of Forest Research 13(4):657–665.
Owens, JN, SJ Morris, and S Misra. 1993. The ultra Parker, GG. 1997. Canopy structure and light environment
structural, histochemical, and biochemical development of an old growth Douglas-fir and Western hemlock
of the post- fertilization megagametophye and the forest. Northwest Science 7(4):261–270.
zygotic embryos of Pseudotsuga menziesii. Canadian Paterson, JM. 1993. Handling and planting methods
Journal of Forest Research 23:816–827. influence field performance of red pine ten years after
Owens, JN, JE Webber, SD Ross, and RP Pharis. 1985. planting. Forestry Chronicle 69(5):589–594.
Interac- tion between gibberellin A4/7 and root- Paulson, KM. 1996. Prolonged cold, moisture treatment as
pruning on the reproductive and vegetative processes conifer seeds at controlled moisture content seed.
in Douglas-fir. Science and Technology 24(1) 75–87.
III. Effects on anatomy of shoot elongation and terminal Pausinger, J. 1877. Abies Douglasii. Erdészeti Lapok 16:380–
bud development. Canadian Journal of Forest Research 381. Pavari, A. 1958. La Douglasia verde in Italia. Monti e
15(2): 354-364.
Boschi
Owens JN, JE Webber, SD Ross, and RP Pharis. 1986. 9(7/8):42–58.
Interac- tion between gibberellin A4/7 and root pruning
Pavari, A, and A de Philippis. 1941. La sperimentatzione di
on the reproductive and vegetative processes in Douglas
specie forestali esotiche in Italia. Annales della
fir. Effects on lateral bud development. Canadian
Sperimen- tazione agraria 38:347–432.
Journal of Forest Research 16:211-221.
Pavle, R. 1967. Uredivanje raznodobnih sastojina sumskih
Owston PW, Stein WI. 1974. Pseudotsuga Carr., Douglas-fir,
sjmen- skih baza obicne smrce (Picea abies L.) i
pp. 674–683 in Seeds of Woody Plants in the United States,
jednodbnih sastojina zelene duglazije (Pseudotsuga
CS Schopmeyer, technical coordinator. USDA Forest
menziesii Franco) primjenom neposredne i posredne
Service Agriculture Handbook 450, Washington, DC.
selekcije i meliorativno sanitarnih mjera (with German
Owston, PW, GA Walters, and R Molina. 1992. Selection of summary: Die Einrichtung ungleichaltriger Bestände der
planting stock, inoculation with mycorrhizal fungi, and Samenbezirke der Fichte (Picea abies (L.) Karst.) und
use of direct seeding, pp. 310-327 in Reforestation gleichaltriger Bestände der Douglasie (Pseudotsuga men-
Practices in Southwestern Oregon and Northern ziesii Franco) durch Anwendung der unmittelbaren Auslese
California, ed. SD Hobbs, SD Tesch, PW Owston, RE sowie der verbessernden und sanitären Massnahmen.)
Stewart, JC Tappeiner II., and GE Wells. Forest Research Poslov- no udruzenje sumsko privrednich organizacija
Laboratory, Oregon State University, Corvallis. 463 p. Zagreb.
Page-Dumroese, DS, H Lowenstein, RT Graham, and AE Pearce, ML. 1980. The IUFRO experiments with Douglas-fir
Har- vey. Soil source, seed source, and organic matter in England and Wales, pp. 381–388 in Background Papers
content effects on Douglas-fir seedling growth. Soil and Douglas-fir Provenances. Vol.1: Proceedings IUFRO
Science Society of America Journal 54:229–233. Joint Meeting of Working Parties S2.02–05 Douglas-fir
Pallardy, SG. Physiology of Woody Plants, Third Edition. 2007. Provenances, S2.02–06 Lodgepole Pine Provenances, S2.02–
Academic Press, Burlington, MA. 464 p. 12 Sitka Spruce Provenances, 52.02–14 Abies Provenances.
Panek, JA, and RH Waring. 1995. Carbon isotope variation Vancouver, Canada 1978. British Columbia Ministry of
in Douglas-fir foliage: improving the δ13C–climate Forests, Information Services Branch, Victoria, BC.
relation- ship. Tree Physiology 15:657–663. Pearson, GA. 1931. Forest Types in the Southwest as Determined
References 327
Pharis, RP, and Ferrell, WK. 1966. Differences in drought
by Climate and Soil. USDA Bulletin 247.
Pellon, AE. 1962. Repoblaciones con especies exóticas de
España, especialmente de Vizcaya, para restaurar el
arbolado desaparecido de las especies indígenas, pp.
510–513 in Asamblea Tecnica Forestal 1962. Ministerio
de Agricultura, Direccion General de Montes, Caza y
Pesca fluvial. Madrid 1963.
Pellon, AE. 1966. Pseudotsuga Douglassi. Abeto de
Douglass, pp. 34–37 in Divulgaciones Forestales.
Ediciones de la Caja de Ahorros Viscaina.
Pennell RI. 1988. Sporogenesis in conifers. Advances in Botani-
cal Research 15:179–196.
Peres, AB. 1964. Espécimes mais representativos da Mata do
Buçaco. Direcção–Geral dos Servicos Florestais e
Aquícolas, Lisbon, Portugal. Estudos e Informação No.
205.
Perks, MP, S O’Reilly, E Monaghan, BA Osborne, and DT
Mitchell. 2001. Chlorophyll fluorescence characteristics,
performance, and survival of freshly lifted and cold
stored Douglas-fir seedlings. Annales des Sciences For-
estières 58:225–235.
Perry, TO. 1971. Dormancy of trees in winter. Science 171:29–
36.
Perry, DA, R Molina, MP Amaranthus. 1987. Mycorrhizae,
mycorrhizospheres, and reforestation: current knowledge
and research needs. Canadian Journal of Forest
Research 17(8):929–940.
Person, HH. 1983. The distribution and productivity of fine
roots in a boreal forest. Plant and Soil 71:81–101.
Peterson, DL. 1985. Crown scorch volume and scorch height:
estimates of post-fire tree condition. Canadian Journal
of Forest Research 15:596–598.
Peterson, DL, and MJ Arbaugh. 1989. Estimating postfire
survival of Douglas-fir in the Cascade Range. Canadian
Journal of Forest Research 19: 530-533.
Peterson, EB. 1964. Growth of Douglas-fir planted on Prince
Edward Island. Forestry Chronicle 40(3):332–333.
Peterson, M, and JR Sutherland. 1989. Grey Mould Control
by Seedling Canopy Reduction Through Under-bench
Ventila- tion and Styroblock Aeration. FRDA Report
007. Canada Economic and Regional Development
Agreement. Brit- ish Columbia Ministry of Forests,
Canadian Forestry Service. 16 p.
Peterson, M. 2008. Fusarium species—British Columbia per-
spective on forest seedling production, pp. 109–125 in
National Proceedings: Forest and Conservation Nursery
As- sociations–2007. Proc. RMRS-P-57. USDA Forest
Service, Rocky Mountain Research Station, Fort Collins,
CO.
Petkova, K. 2004. Der Douglasienanbau in Bulgarien. Unpub-
lished manuscript.
Petri, H. 1986. Zum Douglasienanbau in Rheinland–Pfalz.
Allgemeine Forstzeitschrift 41(34/35):859–862.
Petrovi, D. 1951. Foreign Varieties of Forest Trees (Exotics)
in Serbia. Serbian Academy of Science–Monographs–
Book CLXXXII, Institute for Physiology of
Development, Ge- netics and Selection, Book 1, 1951.
Belgrade. Translation OTS 60–21630, NOLIT
Publishing House, Belgrade 1963.
Pfeifer, A. 1988. Douglas-fir Provenance Research in
Ireland. A Summary of Results. Paper to the European
Commission ad hoc Committee on Seeds and
Propagation Materials, May 1988, Brussels.
Pfister, RD, BL Kovalchik, SF Arno, and RC Presby.
1977. Forest habitat types of Montana. USDA Forest
Service, General Technical Report INT-34.
Intermountain Forest and Range Experiment Station,
Ogden, UT. 174 p.
Pharis, RP. 1991. Physiology of gibberellins in relation to
floral initiation and early floral differentiation, pp. 166–
178 in Gibberellins, ed. N Takahashi, BO Phinney, and J
MacMil- lan. Springer-Verlag, NY.
und Jagdzeitung 138 (9):198–205.
resistance between coastal and inland sources of Pintarić, K. 1989. Proučavanje prirašćivanja IUFRO-
Douglas- fir. Canadian Journal of Botany 44:1651– duglazije različitih provenijencija na oglednoj plohi
1659. "Crnalokva" Bosanska Gradiška, Šumarskilist 9-
Pharis, RP, and RW King. 1985. Gibberellins and 10/89:397-414.
reproduc- tive development in seed plants, Annual
Review of Plant Physiology 36: 517–568.
Pharis, RP, and SD Ross. 1976. Gibberellins: their potential
uses in forestry. Outlook on Agriculture 9:82-87.
Pharis, RP, and SD Ross. 1986a. Flowering of Pinaceae
family conifers with gibberellin A4/7 mixture: how
to accom- plish it, mechanisms, and integration with
early progeny testing, pp. 171–179 in Proceedings,
Conifer Tree Seed in the Inland Mountain West,
General Technical Report INT-203. USDA Forest
Service, Ogden, UT.
Pharis, RP, and SD Ross. 1986b. Hormonal promotion
of flowering in Pinaceae family conifers, pp. 269–
286 in Handbook of Flowering, Vol. 5. Ed. A
Halevy. CRC Press, Boca Raton, FL.
Pharis, RP, Evans, LT, King, RW, and Mander, LN.
1989. Gibberellins and flowering in higher plants:
differing structures yield highly specific effects,
pp. 29–42 in Plant reproduction: from floral
induction to pollination, ed. E. Lord and G. Bernier.
American Society of Plant Physiologists,
Rockville, MD.
Pharis, RP, SD Ross, and EE McMullan. 1980.
Promotion of flowering in the Pinaceae by
gibberellins. III Seedlings of Douglas-fir.
Physiologia Plantarum 50(2):119–126.
Pharis, RP, JE Webber, and SD Ross. 1987. The
promotion of flowering in forest trees by gibberellin
A4/7 and cultural treatments: a review of the
possible mechanisms. Forest Ecology and
Management 19:65–84.
Pharis, RP, D Tomchuck, FD Beall, RM Rauter, and G
Kiss. 1986. Promotion of flowering in white spruce
(Picea glauca) by gibberellin A4/7, auxin
(naphthaleneacetic acid), and the adjunct cultural
treatments of girdling and Ca(NO3)2
fertilization. Canadian Journal of Forest Research 16:340–345.
Pharis, RP, FC Yeh, and BP Dancik. 1991. Superior
growth potential in trees: What is its basis, and can
it be tested for at an early age? Canadian Journal of
Forest Research 21(3):368–374.
Pharis, RP, R Zhang, IB-J. Jiang, BP Dancik, and FC
Yeh. 1992. Differential efficacy of gibberellins in
flowering and vegetative shoot growth, including
heterosis and inherently rapid growth, pp. 13–27
in Progress in Plant Growth Regulation, ed. CM
Karssen, LC van Loon, and D Vreugdenhil.
Kluwer Academic Publishers, Dordrecht, The
Netherlands.
Philipson, JJ. 1988. Root growth in Sitka spruce and
Douglas-fir transplants: dependence on the shoot
and stored carbo- hydrates. Tree Physiology
4(2):101–108.
Philipson, JJ. 1990. Prospects for enhancing flowering
of coni- fers and broadleaves of potential
silvicultural importance in Britain. Forestry
63(3):223–240.
Philipson, JL, and MP Cootts. 1979. The induction of
root dormancy in Picea sitchensis (Bong) Carr. by
abscisic acid. Journal of Experimental Botany
30(116):371–380.
Phillips, IDJ, and PF Wareing. 1958. Studies in the
dormancy of sycamore. I. Seasonal changes in the
growth-sub- stance content of the shoot. Journal of
Experimental Botany 9:350–364.
Pickford, AE. 1929. Studies of seed dissemination in British
Columbia. Forestry Chronicle 5(4):8–16.
Pintarić, K. 1967. Das Höhenwachstum verschiedener
Herkünfte der Douglasie (Pseudotsuga taxifolia
Britt.) in den ersten Lebensjahren. Allgemeine Forst
328 Douglas-fir: The Genus Pseudotsuga
Pirags, DM. 1990. Douglas-fir provenance tests and seed or- Research Contribution No. 73, College of Forest Resources,
chards in Latvia. Paper 2.245 in Proceedings Joint University of Washington, Seattle.
Meeting of Western Forest Genetics Association and Praeger, EM, DP Fowler, and AC Wilson. 1976. Rates of
IUFRO Working Parties S2.02–05, 06, 12, and 14.
Olympia, Washington, Aug.
20–24, 1990. Weyerhaeuser Company, Centralia, WA.
Pirags, M. 1968. Duglazija un tas introdukeija. (With Engl.
summary “Douglas-fir and its introduction”), pp.53–
67 in Povysenie produktivnosti lesa. Izdatelstvo Zinatne,
Riga, Latvia.
Pirags, M. 1979. Duglasija v Latvijskoj SSR. (Douglas-fir in
the Latvian SSR). Riga.
Pirozynski, KA. 1981. Interactions between fungi and plants
through the ages. Canadian Journal of Botany 59:1824–
1827.
Piškorić, O. 1960. Duglazija kao vrsta ekonomskih
sastojina na degradiranom dijelu Krasa. (with Engl.
summary: Douglas-fir as one of the species
constituting the pro- ductive stands in the degraded
part of the Karst region). Sumarski List 84:383–392.
Plaut, M. 1910. Untersuchungen zur Kenntnis der physiol-
gischen Scheiden bei den Gymnospermen, Equiseten,
und Bryophyten. Jahrbücher für wissenschaftliche
Botanik 47:121–185.
Podhorsky, J, 1927. New information on Douglas fir in
Europe.
Allgemeine Forst U Jagdzeitune: 255–258
Poethig, RS. 1990. Phase change and the regulation of
shoot
morphogenesis in plants. Science 250:923–930
Pojar, J, K Klinka, and DV Meidinger. 1987. Biogeoclimatic
ecosystem classification in British Columbia. Forest
Ecol- ogy Management 22:191–154.
Polansky, B. 1934. Lesnické pestováni drevin
cizokrajnych se zretelem na poméry v CSR I
(Silviculture of exotic species in relation to the
condtition of the CSR, part I). Sborník vzkumnych
ústav zemedelskych 124, Prague.
Pollard, DFW, and CC Ying. 1979. Variation in response to
declining photoperiod among families and stands of
white spruce in Southeastern Ontario. Canadian Journal
of Forest Research 9(4):443–448.
Pons, TL. 1983. Significance of inhibition of seed germina-
tion under the leaf canopy in ash coppice. Plant Cell and
Environment 6:385-392.
Poncelet, J. 1963. Les douglas de Mésy, du Hohwald et de
Quinault. Bulletin de la Société Royale de Botanique de
Bel- gique 70(12):567–570.
Poole, BR. Importance of care of stock between the nursery and
the planting site. New Zealand Forest Products, Ltd. 7 p.
Portier, PJ. 1918. Les Symbiotes. Available online at
http://hdl. handle.net/2027/mdp.39015011399261
Potter, LD, and J Rowley. 1960. Pollen rain and
vegetation, San Augustin Plains, New Mexico.
Botanical Gazette 122(1):1–25.
Powell, LE. 1982. Shoot growth in woody plants and
possible participation of abscisic acid, pp. 363–372 in
Plant Growth Substances, ed. PF Wareing, Academic
Press, London.
Powell, LE. 1987a. Hormonal aspects of bud and seed dor-
mancy in temperate-zone woody plants. HortScience
22(5):845–850.
Powell, LE. 1987b. The hormonal control of bud and seed
dor- mancy in woody plants, pp. 539–552 in Plant
Hormones and Their Role in Plant Growth and
Development, ed. PJ Davies. Martinus Nijhoff Publ.,
Dordrecht, The Netherlands.
Powers, RF. 1992. Concluding remarks, pp. 282–283 in
Sustain- ing and Improving Nutrition and Growth of
Western Forests. Proc. Symposium on Forest Fertilization,
Seattle, Washington, February 1991, ed. HN Chappell,
GF Weetman, and RE Miller. Institute of Forest
evolution in conifers (Pinaceae). Evolution 30(4):637–649. Radwan, MA, GL Crouch, and HS Ward. 1971. Nursery fer-
Prat, D. 1995. Mating system in a clonal Douglas fir tilization of Douglas-fir seedlings with different forms of ni-
(Pseu- dotsuga menziesii (Mirb.) Franco) seed
orchard. II. Effec- tive pollen dispersal. Annals of
Forest Science 52:213–222.
Preece, SJ, Jr. 1950. Floristic and Ecological Features of
the Raft River Mountains of Northwestern Utah. MS
thesis, Univer- sity of Utah, Salt Lake City.
Price, RA, J Olson-Stojkovich, and JM Lowenstein.
1987. Rela- tionships among the genera of
Pinaceae: an immunologial comparison. Systemic
Botany 12:91–97.
Price, RA. 1989. The genera of Pinaceae in the southeastern
United States. Journal of the Arnold Arboretum 70:247–305.
Priestley, CA. 1970. Carbohydrate storage and
utilization, pp. 113–127 in Physiology of Tree
Crops, ed. LC Luckwill and CV Cutting, Academic
Press, New York. 380 p.
Prior, KW. 1959. Wind damage to exotic forests in Canterbury.
New Zealand Journal of Forestry 8(1):56–68.
Prior, KW, RW Washbounn, and RM Priest. 1963.
Afforesta- tion in the Otago Land District. New
Zealand Journal of Forestry 8(5):707–727.
Puchert, H. 1954. Das Ergebnis der Anbauversuche mit
fremdlän- dischen Baumarten in Braunschweig. Eine
Untersuchung über die Erfolge des Ausländeranbaus
der ehemaligen Braun- schweigischen
Versuchsanstalt aus den Jahren 1880–1910.
Universitaet Göttingen, Inaugural Dissertation.
Puritch, GS. 1972. Cone Production in Conifers: A
Review of the Literature and Evaluation of Research
Needs. Pacific Forest Research Center, Department
of Environment Informa- tion Report BC-X-65,
Canadian Forestry Service, Victoria, BC. 94 p.
Puritch, GS, EE McMullin, MD Meager, and CS
Simmons. 1979. Hormonal enhancement of cone
production in Douglas-fir grafts and seedlings.
Canadian Journal of Forest Research 9:193–200.
Pirozynski, KA. 1981. Interactions between fungi and
plants through the ages. Canadian Journal of Botany
59(10):1824– 1827.
Rack, K. 1974. Frostanalyse in einer Douglasienkultur
unter Berucksichtigung von Phomopsis pseudotsugae.
Allgemeine Forst- und Jagdzeitung 145(8):154–162.
Racz, J, and J Kleinschmit. 1978. Standörtliche
Grundlagen für Anbauempfehlungen von
Douglasien–Herkünften and erste Ergebnisse der
Prüfung von Bestandesbeern- tungen aus den USA.
Allgemeine Forst– und Jagdzeitung 149(6/7):101–
113.
Radcliffe, DN. 1951. Forest seed and plantation insects,
with special reference to Barbara colfaxiana Kearf. in
Douglas-fir cones. Canadian Department of
Agriculture, Forest Biol- ogy Division Science
Service, Forest Biology Laboratory, Victoria, BC.
44 p.
Radcliffe, DN. 1952. An appraisal of seed damage by
the Douglas-fir cone moth, in British Columbia.
Forest Chron- icle 28:19–24.
Radulovi, S. 1960. Duglazija u swetlosti podataka
njenog razvoja na Avali (Douglas-fir in light of
its development on Mt. Avala). Sumarstvo 13
(9/10):415–424.
Radwan, NA. 1963. Protecting Forest Trees and Their
Seeds From Wild Mammals. USDA Forest Service
Research Paper PNW 6, Pacific Northwest Forest
and Range Experiment Station, Portland, OR.
Radwan, NA. 1971. Factors affecting endrin content of
endrin- coated Douglas-fir seed. Northwest Science
45(3) 188–192.
Radwan, MA, and HW Anderson. 1970. Storing
Endrin-coated and Endrin-Impregnated Douglas-
fir Seed. USDA Forest Service Research Paper
PNW–95, Pacific Northwest For- est and Range
Experiment Station, Portland, OR. 6 p.
References 329
Symposium of the British Mycological Society Held at Bath
trogen. USDA Forest Service Research Paper PNW- University 11–15 April 1983, ed. DH Jennings and ADM
113, Pacific Northwest Forest and Range Experiment Rayner.
Station, Portland, OR.
Radwan, MA, JS Shumway, and DS DeBell. 1979. Effects of
Manganese and Manganese-Nitrogen Applications on Growth
and Nutrition of Douglas-fir Seedlings. USDA Forest
Service Research Paper PNW-265, Pacific Northwest
Forest and Range Experiment Station, Portland, OR.
Rafn, J. 1915. The testing of forest seeds during 25 years, 1887–
1912.
Langkjaers Bogtrykkeri, Copenhagen. 91 p.
Randall WK. 1996. Forest tree seed zones for western Oregon.
Oregon Department of Forestry, Salem, OR.
Randall, WK, and P Berrang. 2002. Washington tree seed
transfer zones. Washington Department of Natural
Resources, Olympia, WA.
Rannert, H. 1958. Zur Inventur der fremdländischen
Baumarten in Österreich. Centralblatt für das gesamte
Forstwesen 75(3/5):284–297.
Rannert, H. 1959a. Bericht über die in Vorarlberg vorkom-
menden fremdländischen Baumarten. Centralblatt für
das gesamte Forstwesen 76(3):162–168.
Rannert, H. 1959b. Bericht über die in Salzburg vorkom-
menden fremdländischen Baumarten. Centralblatt für
das gesamte Forstwesen 77(2):107–115.
Rannert, H. 1960. Bericht über die in Burgenland vorkom-
menden fremdländischen Baumarten. Centralblatt für
das gesamte Forstwesen 77(3):169–182.
Rannert, H. 1973. Aussagen über den Douglasienanbau in
Ös- terreich, pp. 5–9 Rundschreiben I/1073,
Arbeitsgemeinschaft für Auwaldwirtschaft und
Flurholzanbau, Wien–Hadersdorf, Austria.
Rannert, H. 1979. Über den Anbau fremdländischer
Baumarten in Österreich von Cieslars ersten
Anbauversuchen bis zum derzeitigen Stand. Centralblatt
für das gesamte Forstwesen 96(2):86–120.
Rappaport, NG. 1988. Lacinipolia patalis Grote
(Lepidoptera: Noctuidae) infesting Douglas-fir cones: A
new host re- cord. The Canadian Entomologist
(November):1033–1034.
Rappaport, NG, and WJA Volney. 1986. Resource utilization
by insects colonizing Douglas-fir (Pseudotsuga
menziesii (Mirb) Franco) cones, pp. 157–166,
Proceedings of the 2nd Conference on Cone and Seed
Insects Working Party S2.07–01. Briançon, France,
September 3–5, 1986, ed. A Roques, Station de Zoologie
Forestière, Olivet, France.
Rappaport, NG, and WJA Volney. 1989. Effect of pest species
exclusions on Douglas-fir (Pseudotsuga menziesii (Mirb.)
Franco) cone and seed insects: implications for pest man-
agement, pp. 142–150 in Proceedings of the 3rd Cone and
Seed Insects Working Party Conference, held in Victoria,
Brit- ish Columbia, Canada, on 26–30 June 1988, Pacific
Forestry Centre, Forestry Canada, Victoria, BC.
Rappaport, N, S Mori, and A Roques. 1993. Estimating effect
of Megastigmus spermotrophus (Hymenoptera: Torymidae)
on Douglas-fir seed production: the new paradigm.
Journal of Economic Entomology 86(3):845
Rau, HM. 1985. Der Douglasienprovenienzversuch von
1958
in Hessen. Allgemeine Forst– und Jagdzeitung 156(4):72–
79.
Rau, HM. 1987. Comparative research with German and
American Douglas-fir provenances in Proceedings
IUFRO Working Party S2.02–05 Meeting on Breeding
Strategy of Douglas-fir as an Introduced Species.
Vienna, Austria, June 10–14, 1985. FBVA Berichte No.
21:85–103. Schriftenreihe der Forstlichen
Bundesversuchsanstalt Vienna, Austria.
Read, DJ. 1984. The structure and function of the vegetative
mycelium of mycorrhizal roots, pp. 215–240 in The
Ecol- ogy and Physiology of the Fungal Mycelium:
populations of Douglas-fir (Pseudotsuga menziesii var.
British Mycological Society, Cambridge University glauca). USDA Forest Service Research Note INT-255,
Press, Cambridge. Intermountain Forest and Range Experiment Station,
Read, DJ. 1991. Mycorrhizas in ecosystems. Experientia Ogden, UT. 7 p.
47(4):376–391. Rehfeldt, GE. 1979b. Ecological adaptations in Douglas-fir
Read, RA, and JA Sprackling. 1976. Douglas-fir in
Eastern Ne- braska. A Provenance Study. USDA
Forest Service, Research Paper RM–178, Fort
Collins, CO.
Reardon, RC, and L Barrett. 1984. Effects of treating
west- ern spruce budworm populations on grand
fir and Douglas-fir with acephate, carbofuran,
dimethoate, oxydemeton-methyl, and
methamidophos. Forest Ecol- ogy and
Management 8(1):1–10.
Reardon, RC, LJ Barrett, TW Koerber, LE Stipe, and JE
Dewey. 1985. Implantation and injection of
systemics to surpress seed and cone insects in
Douglas-fir in Montana. The Canadian
Entomologist 117:961–969.
Reck, SG. 1978. Height growth and frost resistance in
Douglas- fir provenances tested in the northern part
of Germany, pp. 175–188 in Proceedings of the
IUFR joint meeting of working parties S2–2–12 Sitka
spruce provenances, S2–02–06 Lodgepole pine
provenances, S2–02–12 Sitka spruce prov- enances,
S2–02–14 Abie provenances. Volume one. Vancou-
ver, Canada 1978, British Columbia Ministry of
Forests, Information Services Branch, Victoria, BC.
Redfern, DB. 1978. Infection by Armillaria mellea and
some factors affecting host resistance and the
severity of the disease. Forestry 51(2):121–135.
Rediske, JH. 1961. Maturation of Douglas-fir seed. A
biochemi- cal study. Forest Science 7:32–60.
Rediske, JH. 1969. Effects of picking date on Douglas-fir seed
quality. Forest Science 15(4):404–409.
Rediske, JH, and KR Shea. 1965. Loss of Douglas-fir seed
viability during cone storage. Forest Science 11:463–472.
Reed, KL, JS Shumway, RB Walker, and CS Bledsoe.
1983. Evaluation of the interation of two
environmental factors affecting Douglas-fir
seedlings growth: light and nitrogen. Forest Science
29:193–203.
Reeve, DR, and A Crozier. 1975. Gibberellin bioassays,
pp. 35–64 in Gibberellins and Plant Growth, ed.
HN Krish- namoorthy, John Wiley, New Delhi.
Refn, J. 1966. Traemålings och kulturopgave.
Praktikantarbete på Rössjöholm 1965–66.
Materialet tillhandahållet av jägm Nils Nannestad
Rössjöholms Säteri. Original not seen, cited from
Lemoine and Wirten 1988.
Regal, PJ. 1982. Pollination by wind and animals: Ecology
of geographic patterns. Annual Review of Ecology and
Systemat- ics 13:497–524. DOI:
10.1146/annurev.es.13.110182.002433.
Rego, V, and LR Alvares. 1988. Programa de Acção
Florestal: Tras–os–Montes–um ano de actividade.
Comunicação ao Simposio sobre «A Floresta e o
Ordenamento da Espaço de Montanha». Vila Real
Portugal.
Rehfeldt, GE. 1974a. Local differentiation of
populations of Rocky Mountain Douglas-fir.
Canadian Journal of Forest Research 4(3):399–
406.
Rehfeldt, GE. 1974b. Genetic Variation of Douglas-fir in
the Northern Rocky Mountains. USDA, Forest
Service, Research Note INT–184, Ogden, UT.
Rehfeldt, GE. 1977. Growth and cold hardiness of
intervarietal hybrids of Douglas-fir. Theoretical
and Applied Genetics 50(1):3–15.
Rehfeldt, GE. 1978. Genetic differentiation of Douglas-
fir populations from the northern Rocky Mountains.
Ecol- ogy 59(6):1264–1270.
Rehfeldt, GE. 1979a. Patterns of first–year growth in
330 Douglas-fir: The Genus Pseudotsuga
(Pseudotsuga menziesii var. glauca) populations. I. as affected by light, temperature, and gibberellic acid. Forest
North Idaho and northeast Washington. Heredity
43(3):383–397.
Rehfeldt, GE. 1979c. Variation in cold hardiness among
populations of Pseudotsuga menziesii var. glauca. USDA
Forest Service Research Note INT-233, Intermountain
Forest and Range Experiment Station, Ogden, UT. 11 p.
Rehfeldt, GE. 1982. Ecological Adaptations in Douglas-fir
Popula- tions II. Western Montana. USDA, Forest Service,
Research Paper INT–295.
Rehfeldt, GE. 1983a. Ecological adaptations in Douglas-fir
(Pseudotsuga menziesii var. glauca) populations. III.
Central Idaho. Canadian Journal of Forest Research
13(4):626–632.
Rehfeldt, GE. 1983b. Seed transfer guidelines for Douglas-fir in
central Idaho. USDA Forest Service Research Note INT-
337. Intermountain Forest and Range Experiment
Station, Ogden, UT. 4 p.
Rehfeldt, GE. 1986a. Development and verification of
models of freezing tolerance for Douglas-fir populations
in the Inland Northwest. USDA Forest Service Research
Paper INT-369, Intermountain Forest and Range
Experiment Station, Ogden, UT. 5 p.
Rehfeldt, GE. 1986b. Performance of Douglas-fir hybrids after
10 years of field investigations. USDA Forest Service
Research Note INT-355. Intermountain Forest and
Range Experi- ment Station, Ogden, UT. 2p.
Rehfeldt, GE. 1988. Ecological adaptations in Douglas-fir
(Pseudotsuga menziesii var. glauca) IV. Montana and
Idaho near the Continental Divide. Western Journal of
Applied Forestry 3(4):101–105.
Rehfeldt, GE. 1989. Ecological adaptations in Douglas-fir
(Pseudotsuga menziesii var. glauca): a synthesis. Forest
Ecology and Management 28:203–215.
Rehfeldt, GE. 1991. The genetic resource of Douglas-fir in
the interior Northwest, pp.53–62 in Interior Douglas-fir:
The Species and Its Management, ed, DM Baumgartner
and JE Lotan. Cooperative Extension, Washington State
University, Pullman.
Reichenau, V. 1911. Ausländische Holzarten in den Forsten
des Regierungsbezirks Danzig. Mitteilungen der
Deutschen Dendrologischen Gesellschaft 20:109–114.
Reinke, J. 1872. Andeutungen über den Bau der Wurzel von
Pinus Pinea. Bot. Zeit. 30:49–54.
Research Institute for Forestry and Landscape Planning
Wa- geningen. Annual Report 1981:44–45.
Reukema, DL. 1961. Seed Production of Douglas-fir
Increased by Thinning. USDA Forest Service, Pacific
Northwest Forest and Range Experimental Station
Research Note 10, Portland, OR. 5 p.
Reukema, DL. 1964. Some effects of freeze injury on
develop- ment of Douglas-fir. Northwest Science
38(1):14–17.
Reukema, J. 1982. Seed fall in a young-growth Douglas-
fir stand 1950–1978. Canadian Journal of Forest
Research 12:249–254.
Reuss. 1885. Zur Frage über die Anbauwürdigkeit auslän-
discher Holzarten für unsere Forsten. Zeitschrift für
Forst und Jagdwesen 17(5):249–271.
Revel, J. 1967. Failure of Douglas-fir Seed Crop in the
Prince George Forest District. British Columbia Forest
Service Forest Research Review 99–100.
Revell, DH. 1978. The site limitations of Douglas-fir, pp.
173–183 in A review of Douglas-fir in New Zealand.
FRI Symposium No. 15. ed. RN James and EH Bunn.
New Zealand Forest Service, Forest Research Institute,
Rotorua.
Reynolds, ERC. 1975. Tree rootlets and their distribution, pp.
163–177 in The Development and Function of Roots, ed.
JG Torrey and DT Clarkson, Academic Press, London.
618 p.
Richardson, SD. 1959. Germination of Douglas-fir seeds
Science 5:174–181. Range Experimental Station, Portland, OR. 17 p.
Richardson, HH, and H Roth. 1968. Hydrocyanic acid Robak, H. 1968. Overvintring og produksjonssikkerhet hos Douglas
and other fumigants for control of larvae of i planteskolene i Vest–Agder og Vestlandsfylkene, bedømt et-
Plemeliella abietina and Megastigmus sp. in
imported spruce seed. Journal of Economic
Entomology 61(1):214–216.
Riebeling, R. 1979. Standortkundliche, ertragskundliche
und betriebswirtschaftliche Überlegungen zum
Anbau schnell- wachsender Baumarten,
insbesondere zur Douglasie. Der Forst–und
Holzwirt 34(9):177–184.
Rietveld, WJ. 1989. Transplanting stress in bare root
conifer seedlings: Its development and progression
to establish- ment. Northern Journal of Applied
Forestry 6:99–106.
Ritchie, GA. 1982. Carbohydrate reserves and root
growth potential in Douglas-fir seedlings before and
after cold storage. Canadian Journal of Forest
Research 12:905–912.
Ritchie, GA. 1984. Effect of freezer storage on bud
dormancy release in Douglas-fir seedlings. Canadian
Journal of Forest Research 14:186–190.
Ritchie, GA. 1985. Root growth potential principles, proce-
dures and predictive ability, pp. 93–105 in
Proceedings: Evaluating seedling quality: principles,
procedures, and pre- dictive abilities of major tests.
Workshop held October 16–18, 1984, ed. ML Duryea.
Forest Research Laboratory, Oregon State University,
Corvallis.
Ritchie, GA. 1986. Relationships among bud dormancy
status, cold hardiness, and stress resistance in 2+0
Douglas-fir. New Forests 1(1):29–42.
Ritchie, GA. 1987. Some effects of cold storage on seedling
physiology. Tree Planters’ Notes 38(2):11–15.
Ritchie, GA. 1988. Root growth potentials, principles,
proce- dures, and predictive ability, pp. 93–104 in
Informational Package for Alternative Root Growth
Capacity Testing Systems, LF Ebell, compiler,
Ministry of Forests Lands, British Columbia,
Victoria.
Ritchie, GA. 1990. A rapid method for detecting cold
injury in conifer seedling root systems. Canadian
Journal of Forest Research 20:26–30.
Ritchie, GA, and JR Dunlap 1980. Root growth
potential: Its development and expression in forest
tree seedlings. New Zealand Journal of Forest
Science 10(1):218–248.
Ritchie, GA, and JW Keeley. 1994. Maturation in
Douglas- fir: I. Changes in stem, branch, and
foliage characteris- tics associated with ontogenetic
aging. Tree Physiology 14:1245–1259.
Ritchie, GA, and TD Landis. 2006. Seedling quality
tests: root electrolyte leakage. Forest Nursery Notes
Winter 2006. USDA Forest Service State and
Private Forestry Pacific Northwest Region R6–CP-
TP0805, pp. 1–10.
Ritchie GA, SD Duke, and R Timmis. 1994.
Maturation in Douglas-fir: II. Maturation
characteristics of genetically matched Douglas-fir
seedlings, rooted cuttings and tissue culture
plantlets during and after 5 years of field growth.
Tree Physiology 14(11):1261–75.
Ritchie, GA, and Y Tanaka. 1990. Root growth potential
and the target seedling, pp. 37–51 in Proceedings,
Western For- est Nursery Association, August 13-
17, 1990, Roseburg, OR, ed. R Rose, SJ Campbell,
and T Landis. General Technical Report RM-200,
USDA Forest Service, Rocky Mountain Forest and
Range Experiment Station, Fort Collins, CO.
Available at:
http://www.fcanet.org/proceedings/1990/ ritchie.pdf
Roadwan, MA, GL Crooch, and WD Ellis. 1970.
Impregnating and Rooting with Endrin to Protect
Douglas-fir Seed from Rodents. USDA Forest
Service Research paper PNW–94, PNW Forest and
References 331
Allelopathic effects of liter on the growth and coloniza-
ter såningene 1930–66 (with English summary:
Hibernation and plant yield of Douglas-fir in the nurseries
of Vest–Agder and the West–Norwegian counties. A
judgement based on the sowings 1930–66). Norske
Skogplanteskoler, Kontoret for Skogkultur, Årsskrift
1967:35–53.
Robbins, J. 1986. The Deadly Irony of Endrin. Sierra 66(6):30–32.
Roberts DR, P Toivonen, and SM McInnis SM. 1991.
Discrete proteins associated with overwintering of
interior spruce and Douglas-fir seedlings. Canadian
Journal of Botany 69:437–441.
Robinson, BA. 1963. Variations in Seed Characteristics of
Douglas- fir in British Columbia. BSF thesis, Faculty of
Forestry, University of British Columbia, Vancouver.
Robinson, LW, and PF Wareing. 1969. Experiments on
juvenile- adult phase change in some woody species. The
New Phytologist 68:67–68.
Robinson, RL. 1914. Some Douglas-fir plantations. Quarterly
Journal of Forestry 8(3):187–190.
Rocuant Trucios, LR. 1967. Análisis de las plantaciones de
pino oregón (Pseudotsuga menziessi Mirb. Franco) en
Chile. Uni- versidad de Concepcion, Escuela de
Agronomia, Depar- tamento de Suelos, Chillan, Chile.
Circular Informativa No. 18.
Rodriguez, EM. 1960. Estudio comparado de la estructura de
maderas de pino oregón (Pseudotsuga menziesii) (Mirb.)
Franco. Revista de Investigaciones Forestales 2(1):89–
99.
Roeke, H. 1982. Cone and Seed Evaluations, Seed Cone. 790.
Ministry of Forests, Tree Seed Centre survey, BC, p. 5.
Roeser, J, Jr. 1926. The importance of seed source and the
possibilities for forest tree breeding. Journal of Forestry
24(1):38–51.
Roeser J Jr. 1942. The influence of climate on seed
production in Douglas-fir. Journal of Forestry 40: 304–
307.
Rohmeder, E. 1954. 23–jährige bayerische Anbauversuche
mit grüner Douglasie verschiedener Herkunft. Der Forst
and Holzwirt 9(9):179–180.
Rohmeder, E. 1956. Professor Münchs Anbauversuch mit
Douglasien verschiedener Herkunft and anderen Nadel-
baumarten im Forstamt Kaiserslautern–Ost 1912–1954.
Silvae Genetica 5(5/6):142–156.
Romberger, JA. 1963. Meristems, Growth, and Development
in Woody Plants. USDA Forest Service Techical
Bulletin No. 1293, Forest Physiology Laboratory, Plant
Industry Station, Beltsville, MD.
Romberger, JA. 1967. Flowering as a Problem in Developmen-
tal Physiology, pp. 22–24 in Proceedings XIV International
Union of Forestry Research Organizations (IUFRO)
Congress, Munich.
Romberger, JA, and RA Gregory. 1974. Analytical
morpho- genesis and the physiology of flowering in
trees, pp. 132–147 in Proceedings of the 3rd North
American Forest Biology Workshop, ed. CPP Reid and
GH Fechner, Colorado State University, Fort Collins.
Roques, A. 1991. Structure, specificity, and evolution of
insect guilds related to cones of conifers in western
Europe, pp. 300–315 in Forest Insect Guilds: Patterns of
Interaction with Host Trees, ed. YN Baranchikov, WJ
Mattson, FP Hain, and TL Payne. USDA Forest Service
General Technical Report NE-153, Northeastern
Experiment Station, Radnor, PA.
Rose, R, and D Haase. 2002. Chlorophyll fluorescence and
variation in tissue cold hardiness in response to
freezing stress in Douglas-fir seedlings. New Forests
23(3):81–96.
Rose, R, and JS Ketchum. 2003. Interaction of initial
seedling diameter, fertilization, and weed control on
Douglas-fir growth over the first four years after
planting. Annals of Forest Science 80:1–11.
Rose, SL, DA Perry, D Pilz, and MM Schoenberger. 1983.
the produc- tion and distribution of cone buds in
tion of mycorrhizal fungi. Journal of Chemical Ecology 9(8) Douglas-fir. Silvae Genetica 38:177–185.
1153–1162. Ross, SD, and RP Pharis. 1976. Promotion of flowering in
Ross, DA, and S Ilnytzky. 1977. The black army the Pinaceae by gibberellins. I. Sexually mature, non-
cutworm, Acte- bia fennica (Tauscher) in British flow-
Columbia. Government of Canada, Department of
the Environment, Information Report No. BC-X-
154, Canadian Forest Service, Pacific Forest
Research Centre, Victoria, BC.
Ross, DW, and GE Daterman. 1995. Efficacy of an
antiaggre- gation pheromone for reducing Douglas-
fir beetle, Den- droctonus pseudotsugae Hopkins
(Coleoptera: Scolytidae), infestation in high risk
stands. The Canadian Entomologist 127:805–811.
Ross, DW, and GE Daterman. 1997. Integrating
pheromone and silvicultural methods for managing
the Douglas-fir beetle, pp. 135–145 in Proceedings:
Integrating cultural tactics into the management of
bark beetle and reforestation pests, Vallombrosa,
Italy, September 1–3, 1996, ed. JC Gré- goire, AM
Liebhold, FM Stephen, KR Day, and SM Salom.
USDA Forest Service General Technical Report
NE-236, Northeastern Forest Experiment Station,
Radnor, PA. Online here: http://iufro-
archive.boku.ac.at/wu70307/
vallombrosa_pictures/valproc/.
Ross, DW, KE Gibson, and GE Daterman. 2006. Using
MCH to Protect Trees and Stands from Douglas-fir
Beetle Infesta- tion. Report FHTET-2001–09, Forest
Health Technology Enterprise Team, USDA Forest
Service, Morgantown, WV.
Ross, DW, GE Daterman, JL Boughton, and TM
Quigley. 2001. Forest Health Restoration in South-
Central Alaska: A Problem Analysis. USDA Forest
Service General Techni- cal Report PNW-GTR-
523, Pacific Northwest Research Station, Portland,
OR.
Ross, SD. 1969. Gross metabolic activity accompanying
the after-ripening of dormant Douglas-fir seeds.
Botanical Gazette 130(4):271–275.
Ross, SD. 1976. Differential flowering responses by
young Douglas-fir grafts and equi-sized seedlings to
gibberellins and auxin. Acta Horticulturae 56:163–
168.
Ross, SD. 1983. Enhancement of shoot elongation in
Douglas- fir by gibberellin A4/7 and its relation to
the hormonal promotion of flowering. Canadian
Journal of Forest Research 13:986–994.
Ross, SD. 1988. Effects of temperature, drought, and
gibberel- lin A417, and timing of treatment, on
flowering in potted Picea engelmannii and Picea
glauca grafts. Canadian Journal of Forest Research
18:163–171.
Ross SD. 1989. Long term cone production and
growth re- sponses to crown management and
gibberellin A4/7 treatment in a young western
hemlock seed orchard. New Forests 3: 235–245.
Ross, SD. 1990. Control of sex expression in potted
Picea engelmannii grafts by gibberellin A4/7 and
the auxin, naphthaleneacetic acid. Canadian Journal
of Forest Research 20(7): 875–879.
Ross, SD. 1991. Effect of heat sums and of heat applied
sepa- rately to shoots on flowering in potted Picea
glauca grafts. Canadian Journal of Forest
Research 21:672–679.
Ross, SD, and RC Bower. 1989. Cost effective
promotion of flowering in a Douglas-fir seed
orchard by girdling and pulsed stem injection of
gibberellin A4/7. Silvae Genetica 38:189–195
Ross, SD, and RC Bower. 1991. Promotion of seed
production in Douglas-fir grafts by girdling plus
gibberellin A4/7 stem injection and effect of
retreatment. New Forests 5:23-34.
Ross, SD, and RC Currell. 1989. Effect of top pruning
branch thinning and gibberellin A4/7 treatment on
332 Douglas-fir: The Genus Pseudotsuga
ering grafts of Douglas-fir. Plant Physiology 36:182– MN. 8 p.
186. Ruetz, WF, R Dimpflmeier, J Kleinschmitt, J Svolba, H Weis-
Ross, SD, and RP Pharis. 1985. Flower induction in crop
trees: different mechanisms and techniques with
special refer- ence to conifers, pp. 383–397 in Attributes
of Trees as Crop Plants, ed. MGR Cannell, JE
Jackson, and JC Gordon. In- stitute of Terrestrial
Ecology, Monkswood Experimental Station, Abbots
Ripton, Huntingdon, UK.
Ross, SD, and RP Pharis. 1987. Control of sex expression in
conifers. Journal of Plant Growth Regulation 6:37–60.
Ross, SD, RP Pharis, and JC Heaman. 1980. Promotion of
cone and seed production in grafted and seedling
Douglas-fir seed orchards by application of gibberellin
A4/7 mixture. Canadian Journal of Forest Research
10(4):464–469.
Ross, SE, RW Skadsen, and RP Pharis. 1976. Progress in the
promotion of early flowering in Douglas-fir by gibberel-
lins. pp. 197–198 in Proceedings of the Fourth North
American Forest Biology Workshop, Syracuse, NY.
Ross, SD, JE Webber, RP Pharis, and JN Owens. 1985.
Interac- tion between gibberellin A4/7 and root-pruning
on the reproductive and vegetative process in Douglas-
fir. I. Effects on flowering. Canadian Journal of Forest
Research 15(2):341–347.
Rossa, ML, and JB Larsen. 1980. Die winterlichen Austrock-
nungsraten verschiedener Herkunfte der Douglasie
(Pseudotsuga menziesii) und deren Abhangigkeit von
der Ausbildung der Cuticula und der Spaltoffnungstiefe.
Allgemeine Forst und Jagdzeitung 151(8):137–148.
[Winter drying rates of different Douglas fir (Pseudotsuga
menziesii) provenances, and their relation to cuticle
thickness and stomata depth.]
Roth, H, and GE Strasser. 1971. Laboratory tests with a
mix- ture of carbon disulfide and carbon tetrachloride
for quarantine control of larvae of Megastigmus Spp.
and Cecidomyiidae in conifer seeds. Journal of
Economic En- tomology 64(4):904–906.
Rothkirch, Fv, and B Struthoff. 1989. Die ältesten
Douglasien in Deutschland–150 Jahre alt. Forst und Holz
44(18):499–500.
Roux, G, A Roques, and F Menu. 1997. Effect of
temperature and photoperiod on diapause development
in a Douglas- fir seed chalcid Megastigmus
spermotrophus. Oecologia 111(2):172–177.
Rowe JS. 1964. Environmental preconditioning, with special
reference to forestry. Ecology 45:399–403.
Rowe, KE, and KK Ching. 1973. Provenance study of
Douglas- fir in the Pacific Northwest region. II. Field
performance at age nine. Silvae Genetica 22(4):115–119.
Rohwer, SA. 1913. Chalcidids injurious to forest-tree seeds.,
pp.
155–161 in Technical Papers on Miscellaneous Forest Insects.
VI. Technical Series No. 20, Part VI. Bureau of Entomol-
ogy, U.S. Department of Agriculture, Washington, DC.
Roy, DF. 1957. Seed spot tests with tetramine-treated seed
in
Northern California. Journal of Forestry 55:442–445.
Roy, DF. 1960. Douglas-fir Seed Dispersal in Northwestern
Cali- fornia. USDA Forest Service Technical Paper No.
49, Pa- cific Southwest Forest and Range Experiment
Station, Berkeley, CA. 22 p.
Rozenberg, P. 1993. Comparaison de la croissance en
hauteur entre 1 et 25 ans de 12 provenances de douglas
(Pseu- dotsuga menziesii (Mirb.) Franco). Annals of
Forest Science 50(4):363–381.
Rudenko, LP, and MA Derzhanovskaya. 1985. (Douglas-
fir in the forests of the Carpathians). Lesnoe
Khozyaistvo 1985, No. 10:69–71. Original not seen,
cited from Forestry Abstracts 47:3965, 1986.
Rudolf, PO. 1952. Low temperature seed storage for western
conifers. Miscellaneous Report 20, USDA Forest
Service, Lake States Forest Experiment Station, St. Paul,
berger, and HM Rau. 1990. The IUFRO Abies initia- tion of pharate adult development in Barbara
procera provenance trial in the Federal Republic of colfaxiana (Kft.) (Lepidoptera: Olethreutidae).
Germany - field results at age 9 and 10 years, pp. Canadian Journal of Zoology 61(10):2305–2306.
2.276–2.289 in Proceed- ings of the Joint Meeting Sahota,TS, A Ibaraki, and SH Farris. 1985. Pharate-
of Western Forest Genetics Assoc. and IUFRO adult
Working Parties S2.02 - 05, 06, 12, and 14 on diapause of Barbara colfaxiana (Kft.): differentiation of
Douglas-fir, Contorta Pine, Sitka Spruce and Abies
Breeding and Genetic Resources held at Olympia,
Washington, USA Aug. 20-24, 1990. Weyerhauser
Co., Tacoma, WA.
Rumpf, G. 1904. Rhizodermis, Hypodermis und Endodermis
der Farnwurzel. Bibliotheca Botanica 13, Heft 62:1–48.
Ruskov, M, and Petrov, P. 1963. (Investigations on
red deer damage in Bulgarian beech forests). In
Bulgarian with German summary:
Untersuchungen über die durch das Rotwild in
den bulgarischen Buchenwäldern verursa- chten
Schäden und Massnahmen zu deren Verhütung.
Vissh Lesotekhnicheski Institut Nauchne Trudove
11:107–120.
Russell, K. 1971. Root rot influences management of
ice-storm damaged timber. State of Washington
DNR Notes, Note No. 3. 3 p.
Ruth, RH, and RA Yoder. 1953. Reducing wind damage
in the forests of the Oregon Coast Range. USDA
Forest Service Research Paper No. 7, Pacific
Northwest Forest and Range Experiment Station,
Portland, OR. 30 p.
Ruth, D. 1980. A guide to insect pests in Douglas-fir seed
orchards. Canadian Forestry Service, BC-X-204.
Pacific Forest Re- search Centre, Canadian Forestry
Service, Victoria, BC.
Ruth, DS, and AF Hedlin. 1974. Temperature treatment
of Douglas-fir seeds to control the seed chalcid
Megastigmus spermotrophus Wachtl. Canadian
Journal of Forest Research 4(4):441–445.
Ryan, MG, and BJ Yoder. 1997. Hydraulic limits to tree
height and tree growth. BioScience 47:235–242.
Ryan MG, D Binkley, and JH Fownes. 1997. Age-
related decline in forest productivity: pattern and
process. Advances in Ecological Research 27:213–
262.
Ryan, MG, N Phillips, and BJ Bond. 2006. Hydraulic
limita- tion hypothesis revisited. Plant, Cell and
Environment 29:367–381
Ryan, KC, DL Peterson, and ED Reinhardt. 1988.
Modeling long-term fire-caused mortality of
Douglas-fir. Forest Science 34(1):190–199.
Rygiewicz, PT, and CS Bledsoe. 1986. Effects of
pretreatment conditions on ammonium and nitrate
uptake by Douglas- fir seedlings. Tree Physiology
1(2):145–150.
Rygiewicz, PT, and CS Bledsoe. 1984. Mycorrhizal
effects on potassium fluxes by Northwest
coniferous seedlings. Plant Physiology 76(4):918–
923.
Ryker, RA. 1975. A survey of factors affecting
regeneration of Rocky Mountain Douglas-fir.
USDA Forest Service, Re- search Paper INT-174,
Intermountain Forest and Range Experiment
Station, Ogden, UT. 19 p.
Ryker, RA, and J Losensky. 1983. Ponderosa pine and
Rocky Mountain Douglas-fir, pp. 53–55 in
Silvicultural Systems for the Major Forest Types of
the United States, RM Burns, tech. comp. USDA
Forest Service, Agriculture Handbook
445. Washington, DC.
Rzedowski, J, and ML Huerta. 1978. Vegetacion de Mexico.
Editorial Limusa, Mexico.
Sahota, TS, and A Ibaraki. 1991. 1- and 2-year
dormancy of the Douglas-fir cone moth Barbara
colfaxiana (Kft.) (Lepi- doptera: Olethreutidae):
Possible relation to individual weights. The
Canadian Entomologist 123 (5):1153–1155.
Sahota, TS, SH Farris, and AI Ibaraki. 1983. Timing of
References 333
Schenk, CA. 1928. Ausländerei und Herkunftsfragen. Der
1- and 2-year dormancy. The Canadian Entomologist Deutsche Forstwirt 10:472–473.
117(7):873–876.
Sahota, TS, DS Ruth, AI Ibaraki, SH Farris, and FG Peet.
1982. Diapause in the pharate adult stage of insect
development. The Canadian Entomologist
114(12):1179–1183.
Sakai, A, and CJ Weiser. 1973. Freezing resistance of
trees in North America with reference to tree regions.
Ecology 54(1):118–126.
Salisbury, FB. 1986. Dormancy terminology. Hortscience
21(2):185–186.
Salisbury, PJ. 1953. Some Aspects of Conifer Seed Micro Flora.
Canada Department of Agriculture Forest Biology
Division Bi-monthly Progress Report 9 (6) 3–4 (Original
not seen).
Salisbury, PJ. 1955. Mould of Stored Douglas-fir Seed in Brit-
ish Columbia Interim Report. Forest Biology Laboratory,
Victoria, BC.
Samish, RM. 1954. Dormancy in woody plants. Annual Review
of Plant Physiology 5:183–204.
Sanders, P. 1978. Phytohormones and bud dormancy, pp.
423–441 in Phytohormones and Related Compounds -
A Comprehensive Treatise. ed. DS Letham, PB
Goodwin, and TJV Higgins. Elsevier/North Holland
Biomedical Press, New York.
Santantonio, D, and RK Hermann. 1985. Standing crop,
pro- duction, and turnover of fine roots on dry,
moderate, and wet sites of mature Douglas-fir in
western Oregon. Annales des Sciences Forestières
42(2):113–142.
Santantonio, D, RK Hermann, and WS Overton. 1977. Root
biomass studies in forest ecosystems. Pedobiologia 17:1–
31.
Sargent, CS. 1884. Report on the Forests of North America
(exclu- sive Mexico). Department of the Interior, Census
Office, Washington, DC.
Sargent, CS. 1898. The Silva of North America. Vol. 12.
Houghton
Mifflin, Boston, MA.
Sarvas, R. 1974. Investigations on the annual cycle of devel-
opment of forest trees. II. Autumn dormancy and winter
dormancy. Communicationes Instituti Forestalis
Fenniae 84:1–101.
Saunders, P. 1978. Phytohormones and bud dormancy, pp.
423–445 in Phytohormones and related compounds - a
com- prehensive treatise, vol. II, ed DS Letham, PB
Goodwin, and TJV Higgins. Elsevier/North-Holland
Biomedical Press, New York.
Scagel, CF, and RG Linderman. 1998. Influence of ectomy-
corrhizal fungal inoculation on growth and root IAA
concentrations of transplanted conifers. Tree Physiology
18(11):739–747.
Scagel, CF, and RG Linderman. 2000. Changes in root IAA
content and growth of bare root conifers treated with
plant growth regulating substances at planting. Journal
of Environmental Horticulture 18(2):99–107.
Scagel, CF, and RG Linderman. 2001. Modification of root
IAA concentrations, tree growth, and survival by
application of plant growth regulating substances to
container grown conifers. New Forests 18:159–186.
Scagel, RK, R Davidson, YA El–Kassaby, and O Sziklai.
1987. Variation of cone scale and seed morphology in
Douglas- fir (Pseudotsuga menziesii), pp. 107–125 in
FBVA Berichte No. 21, Schriftenreihe der Forstlichen.
Bundesversuchsanstalt, Vienna, Austria.
Scheffer, TC, and GG Hedgcock. 1955. Injury to northwestern
forest trees by sulfur dioxide from smelters. USDA Forest
Service Technical Bulletin No. 1117, Washington, DC. 49
p.
Scheifele, M. 1965. Zur Planung und Wahl der Holzarten in
den öffentlichen Waldungen von Baden–Württemberg.
Allgemeine Forst– und Jagdzeitung 136(12):283–292.
Schoenberger, MM, and DA Perry. 1982. The effect of soil
Schenk, CA. 1939. Fremdländische Wald– und Parkbäume. Vol. dis- turbance on growth and ectomycorrhizae of
2. Die Nadelhölzer. Paul Parey, Berlin. Douglas-fir and western hemlock seedlings; a
greenhouse bioassay. Canadian Journal of Forest
Schenk, JA, RH Giles, Jr, and FD Johnson. 1967. Research 12(2) 343–353.
Effects of trunk-injected oxydemetonmethyl on
Douglas-fir cone and seed insects, seedling
production, and mice. Station Paper No. 2, Forest
Wildlife and Range Experiment Station, University
of Idaho, Moscow. 20 p.
Scheumann, W. 1962. Untersuchungen zur
Frostresistenz der Douglasie (Pseudotsuga taxifolia
(Poir.) Britton) and Fichte (Picea abies (1.) Karst.).
Deutsche Akademie der Land-
wirtschaftswissenschaften, Tagungsberichte
53:155–162.
Schippel, H. 1950. Ein Mahnwort an Douglasienwirte. Allge-
meine Forstzeitung 5(49):351.
Schisler, DA, and RG Linderman. 1984. Evidence for
involve- ment of the soil micro biota in the
exclusion of Fusarium from coniferous forest soils.
Canadian Journal of Microbiol- ogy 30(2):142–150.
Schisler, DA, and RG Linderman. 1989. Influence of
humic- rich organic amendments to coniferous
nursery soils on Douglas-fir growth, damping off
and associated soil mi- croorganisms. Soil Biology
and Biochemistry 21(3):403–408.
Schisler, DA, and RG Linderman. 1989. Selective
influence of volatiles purged from coniferous forest
and nursery soils on microbes of a nursery soil. Soil
Biology and Biochemistry 21(3):389–396.
Schlich, W. 1888. The Douglas fir in Scotland. Gardeners
Chron- icle 4:531–532, 568–569, 598–600. Reprinted in
Royal Scottish Arboricultural Society Transactions
12:226–241.
Schmidt, RL. 1957. The silvics and plant geography of
the genus Abies in the coastal forests of British
Columbia. Department of Lands and Forests,
British Columbia Forest Service, Technical
Publication T.46, Victoria, BC. 31 p.
Schmidt, RL. 1960. Factors controlling the distribution
of Douglas fir in coastal British Columbia.
Quarterly Journal of Forestry 54(3):156–160.
Schmidt, RL. 1973. A Provenance Test of Coastal Douglas
Fir in British Columbia. Preliminary Test Results at 6
Years, pp. 68–81 in Proceedings, IUFRO Working
Party on Douglas-Fir Provenances. Göttingen, West
Germany, September 3–5, 1973.
Schmidt, WC. 1969. Seedbed treatments influence
seedling de- velopment in western larch forests.
USDA Forest Service, Research Note INT-93.
Intermountain Forest and Range Experiment
Station, Ogden, UT. 7 p.
Schmidtling, RC. 1974. Fruitfulness in conifers:
Nitrogen carbohydrate, and genetic control, pp.
148–164 in 3rd North American Forest Biology
Workshop Proceedings, Fort Collins, CO.
Schober, R. 1954. Douglasien-Provenienzversuche I.
Allgemeine Forst– und Jagdzeitung 125(5):160–
179.
Schober, R. 1955. Die Ertragsleistung der Nadelhoelzer
in Grossbritannien und Deutschland–verglichen
nach der neuen britischen Ertragstafel von F.C.
Hummel und J. Christie 1953. Forstwissenschaftliche
Centralblatt 74(1/2):36– 59.
Schober, R, and H Meyer. 1955. Douglasien-
Provenienzver- suche II. Allgemeine Forst- und
Jagdzeitung 126:221–243.
Schober, R, J Kleinschmit, J Svolba. 1983. Ergebnisse
des Douglasien-Provenienzversuches von 1958 in
Nordwes- tdeutschland. I. Teil. Allgemeine Forst–
und Jagdzeitung 154(12):209–236.
Schober, R, J Kleinschmit, J Svolba. 1984. Ergebnisse
des Douglasien-Provenienzversuches von 1958 in
Nordwes- tdeutschland. II. Teil. Allgemeine Forst-
und Jagdzeitung 155(2/3):53–80.
334 Douglas-fir: The Genus Pseudotsuga
Schönbach, H. 1953. Beobachtungen an Einzelstamm- Sexton. 1986. Response of Douglas-fir cone gall midge and
Nach- kommenschaften einheimischer Doug- las-fir seed chalcid to host plant genotype, pp. 217–
Douglasienbestande. Zugleich ein Beitrag zur Frage 23,
der natürlichen Formen- mannigfaltigkeit der
Douglasie. Archiv für Forstwesen 2(6):502–531.
Schönbach, H. 1958. Die Zuchtung der Douglasie. P. 309–
367 in Göhre, K. Die Douglasie und ihr Holz. Akademie
Verlag, Berlin.
Schönbach, H. 1959. The variation of frost resistance in
home- grown stands of Douglas-fir. Recent advantages
in botany.
IX. Int. Botanical Congress Montreal, col. II:1604–
1606. From lectures and symposia presented to the
University of Toronto Press.
Schönbach, H, and E Bellmann. 1967. Frostresistenz der
Nach- kommenschaften von Kreuzungen “grüner"and
“blauer” Formen der Douglasie (Pseudotsuga menziesii
(Mirb.) Franco). Archiv für Forstwesen 16(6/9):707–
711.
Schönbach, H, and H Schmiedel. 1973. Verhalten von Ras-
senhybriden der Douglasie (Pseudotsuga menziesii
(Mirb.) Franco) unter Frosteinwirkung und frostfreien
Bedingun- gen. Beiträge für die Forstwirtschaft 7(3):93–
96.
Schonhar, S. 1965. Frosttrocknis bei der Douglasie.
Allgemeine
Forstzeitung 20(4):44–45.
Schowalter, TD. 1984a. Comparison of arthropods
emerging in the spring from Douglas-fir litter between
a mature stand and a seed orchard in western Oregon.
Environ- mental Entomology 13:1253–55.
Schowalter, TD. 1984b. Dispersal of cone and seed insects
to an isolated Douglas-fir tree in western Oregon. The
Canadian Entomologist 116:1437–38.
Schowalter, TD. 1986. Ecological strategies of forest insects:
the need for a community-level approach to
reforestation. New Forests 1:57–66.
Schowalter, TD. 1988. Forest pest management: a synopsis.
Northwest Environmental Journal 4:313–18.
Schowalter, TD. 1995. Canopy arthropod communities in
relation to forest age and alternative harvest practices in
western Oregon. Forest Ecology and Management 78(1–
3, October):115–125.
Schowalter, TD, and MI Haverty. 1989. Influence of host
geno- type on Douglas-fir seed losses to Contarinia
oregonensis (Diptera: Cecidomyiidae) and Megastigmus
spermotrophus (Hymenoptera: Torymidae) in western
Oregon. Environ- mental Entomology 18:94–97.
Schowalter, TD, and JM Sexton. 1989. Inventory monitoring
as a means of assessing insect impact on Douglas-fir
seed production in western Oregon, pp. 151–60, in
Proceedings of the 3rd Cone and Seed Insects Working
Party Conference, Working Party S2.07-01, IUFRO, 26–30
June 1988, Victoria, BC. GE Miller, comp. Forestry
Canada, Pacific Forestry Research Centre, Victoria, BC.
242 p.
Schowalter, TD, and JM Sexton. 1990. Effect of
Leptoglossus occidentalis (Heteroptera: Coreidae) on
seed development of Douglas-fir at different times
during the growing season in western Oregon. Journal
of Economic Entomol- ogy 83(4):1485–86.
Schowalter, TD, and JD Stein. 1987. Influence of
Douglas-fir seedling provenance and proximity to
insect population sources on susceptibility to Lygus
hesperus (Heteroptera: Miridae) in a forest nursery in
Western Oregon. Environ- mental Entomology
16:984–986.
Schowalter, TD, MI Haverty, and TW Koerber. 1985.
Cone and seed insects in Douglas-fir (Pseudotsuga
menziesii (Mirb.) Franco) seed orchards in the western
United States: distribution and relative impact. The
Canadian Entomologist 117:1223–1230.
Schowalter, TD, MI Haverty, SA Dombrosky and J
in Proceedings 2nd Cone and Seed Insects Working Scott, DRM. 1981. The Pacific Northwest region, pp. 447–493
Party Conference S 2.07-01, Station de Zoologie in Regional Silviculture of the United States, ed. JW
Forestiére, INRA, CRF, Olivet, France, ed. A Roques. Barrett. John Wiley, New York.
Schubert, GH. 1954. Viability of various coniferous seeds after
cold storage. Journal of Forestry 52:446–447.
Schubert, GH, and RS Adams. 1971. Reforestation
practices for conifers in California. California
State Division of Forestry, Sacramento. 359 p.
Schultze U, and Raschka, HD. 2002. Douglas fir
provenances for the summer dry east of Austria –
results of Douglas fir provenances trials of the
Institute of Forest Genet- ics FBVA-Wienna,
Forstliche Bundesverzuchsanstalt Wien:126–195.
Schülli, L. 1984. Erfahrungen mit der Douglasie
zwischen Oberrhein und Bodensee. Allegemeine
Forstzeitschrift 39(17/18):436–437.
Schuch, UK, ML Duryea, and LH Fuchigami. 1989a.
Frost hardiness as acquired by Douglas-fir seedlings
in three Pacific Northwest nurseries. Canadian
Journal of Forest Research 19:192-197.
Schuch, UK, ML Duryea, and LH Fuchigami. 1989b.
De- hardening and budburst of Douglas-fir
seedlings raised in three Pacific Northwest
nurseries. Canadian Journal of Forest Research
19:198-203.
Statistical Analysis Systems Institute Inc., 1988. SAS/STAT
Schupp, EW. 1995. Seed-seedling conflicts, habitat
choice, and patterns of plant recruitment, American
Journal of Botany 82(3):399–409.
Schwabe, WW. 1976. Applied aspects of juvenility and some
theoretical considerations. Acta Horticulturae 56:45–
53. Schwager, G. 1979. Beitrag zur Geschichte der
fremdländischen
Baumarten in der Schweiz–unter besonderer Berücksichtigung
der Weymouthsföhre. Diplomarbeit. Eidgenössische
Tech- nische Hochschule, Zuerich.
Schwappach, A. 1891. Denkschrift betreffend die
Ergebnisse der in den Jahren 1881–1890 in den
Preussischen Staatsfor- sten ausgeführten
Anbauversuche mit fremdländischen Holzarten.
Zeitschrift für Forst- und Jagdwesen 23(1):18–34,
(2):81–102, (3):148–164.
Schwappach, A. 1901. Die Ergebnisse der in den
Preussischen Staatsforsten ausgefuehrten
Anbauversuche mit frem- dländischen Holzarten.
Zeitschrift für Forst- und Jagdwesen 33(3):137–169,
(4):195–225, (5):261–292.
Schwappach, A. 1911. Die weitere Entwicklung der
Versuche mit fremdländischen Holzarten in
Preussen. Zeitschrift für Forst- und Jagdwesen
43(8):591–611, (10):757–782. (Same paper published
in Mitteilungen der Deutschen Dendrolo- gischen
Gesellschaft 20:3–37, 1911).
Schwappach, A. 1920. Beiträge zur Kenntnis der
Wachs- tumsleistungen von Pseudotsuga Douglasii.
Mitteilungen der Deutschen Dendrologischen
Gesellschaft 29:262–269.
Schwarz, H. 1932a. Die Standorts– und
Bestandesverhältnisse der Kuestendouglasie im
Optimum ihres Verbreitungsge- bietes in
Nordamerika. Allgemeine Forst– und Jagdzeitung
108:196–203.
Schwarz, H. 1932b. Über die klimatischen
Möglichkeiten des Anbaues der Küstendouglasie
(Pseudotsuga taxifolia var. viridis) in Österreich.
Centralblatt für das gesamte Forstwesen 58(1):11–
20.
Schwarz, H. 1933. Wuchsgebiete der Küstendouglasie
in Eu- rope. Centralblatt für das gesamte Forstwesen
59(4):110–125.
Schwendemann, AB, G Wang, ML Mertz, RT
Mcwilliams, SL Thatcher, and JM Osborn. 2007.
Aerodynamics of saccate pollen and its implications
for wind pollination. American Journal of Botany
94(8):1371–1381.
References 335
Canada R&D Update (July), Pacific Forestry Centre,
Scott, F. 1931. The place of Douglas fir in Scottish forestry. Victoria, BC.
Forestry 5(1):14–20. Shields, WJ, and SD Hobbs. 1979. Soil nutrient levels and pH
Scott, W, R Meade, R Leon, D Hyink, and R Miller. 1998. associated with Armillaria mellea on conifers in Northern
Plant- ing density and tree size relations in coast
Douglas-fir. Canadian Journal of Forest Research
28:74-78.
Sedgley, M, and AR Griffin. 1989. Sexual Reproduction of
Tree Crops. Academic Press, London.
Seeley, SD. 1994. Dormancy–the black box. HortScience
29(11):1248–1249
Seeley, SD. 1997. Quantification of endodromancy in
seeds of woody plants. HortScience 32(4) 615–617.
Seibert, P. 1951. Die Douglasie im Stadtwald Freiburg. Allge-
meine Forst– und Jagdzeitung 122(6/7):158–161.
Seibert, P. 1979. Die Vegetationskarte des Gebietes von El
Bolson, Prov. Rio Negro, und ihre Anwendung in der
Landnutzung- splanung. Bonner Geographische
Abhandlungen No. 62.
Seidel, S, and D Keyes. 1983. Can We Delay a Greenhouse
Warm- ing? US Environmental Protection Agency,
Washington, DC, 192 p.
Semrau, KT. 1957. Emission of fluorides from industrial
processes–a review. Journal of the Air Pollution Control
Association 7:92–108.
Sharpe, AL, and WL Mason. 1992. Some methods of cold
storage can seriously affect root growth potential and
root moisture content and subsequent forest
performance of Sitka spruce and Douglas fir
transplants. Forestry 65:463–472.
Shaw, EW. 1953. Effects of Tetramine Used for Rodent
Control in Direct Seeding of Douglas-fir. USDA Forest
Service Pacific Northwest Forest and Range Station
Research Note No. 89, Portland, OR. 5 p.
Shaw, TM, JA Moore, and JD Marshall. 1998. Root
chemistry of Douglas-fir seedlings grown under different
nitrogen and potassium regimes. Canadian Journal of
Forest Research 28:1566–1573.
Shaw, DC, GM Filip, A Kanaskie, DA Maguire, and WA Littke.
2011. Managing an epidemic of Swiss needle cast in the
Douglas-fir region of Oregon: The role of the Swiss Needle
Cast Cooperative. Journal of Forestry 109 (March):109–
119.
Shea, KR. 1960. Mold Fungi on Forest Tree Seed. Weyerhaeuser
Co. Forestry Research Note No. 3, Centralia, WA. 10 p.
Shea, KR. 1961. Field Survival of Thiram-treated Douglas-fir
Seed. Weyerhaeuser Co. Forest Research Note No. 38. 8 p.
Shea, KR. 1962. Decay in a freeze-damaged Douglas-fir
plan- tation. Weyerhaeuser Timber Co., Forest
Research Note No. 47, 10 p.
Shearer, RC. 1984. Influence of insects on Douglas-fir, Pseu-
dotsuga menziesii (Mirb.) Franco, and western larch,
Larix occidentalis Nutt., cone and seed production in
western Montana, pp. 112–121 in Proceedings of the
IUFRO Cone and Seed Insects Working Party
Conference. Working Party S2.07–01. July 31–August 6,
1983, Athens, GA, ed. HO Yates,
III. USDA Forest Service Southeastern Forest
Experiment
Station, Asheville, NC. 214 p.
Shearer, RC. 1985. Cone production on Douglas-fir and
west- ern larch in Montana, pp. 63–67 in Proceedings–
Conifer Tree Seed in the Inland Mountain West
Symposium. comp. RC Shearer. USDA Forest Service
GT Report INT-203, Intermountain Research Station,
Ogden, UT. 289 p.
Shearer, RC, and D Tackle. 1960. Effect of Hydrogen
Peroxide on Germination in the Western Conifers.
USDA Forest Service Research Note INT-80,
Intermountain Forest and Range Experiment Station,
Ogden, UT. 4 p.
Shepherd, RF, TG Gray, and TF Maher. 1993.
Management of the black army cutworm. Forestry
improve- ment program in coastal Douglas-fir. Journal
Idaho. Canadian Journal of Forest Research 9:45–48. Forestry 77(2):78–83.
Shirazi, AM, and PS Muir. 1998. In vitro effect of Silen, RR, and DL Olson. 1992. A pioneer exotic tree search
formalde- hyde on Douglas fir pollen. Plant, Cell & for the Douglas-fir region. General Technical Report
Environment 2(3):341–346. PNW- GTR-298. USDA Forest Service, Pacific
Northwest Re- search Station, Portland, OR. 44 p.
Shortt, RL, BJ Hawkins, and JH Woods. 1996.
Inbreeding ef- fects on the spring frost hardiness of Silen, RR, and C Osterhaus. 1979. Reduction of genetic base
coastal Douglas-fir. Canadian Journal of Forest by sizing of bulked Douglas-fir seed lots. Tree Planters'
Research 26(6):1049–1054. Notes 30:24–30.
Sidle, RC, and DM Drlica. 1981. Soil compaction from
logging with a low ground pressure skidder in the
Oregon Coast Ranges. Soil Science of America
Journal 45(6):1219–1224.
Siggins, HW. 1928. What’s New in Agriculture.
Yearbook of
Agriculture 1927:586–587. USDA, Washington, DC.
Sika, A. 1979. Distribution of Douglas-fir in forests of
the CSR.
Communications Instituti Forestalis Cechosloveniae
11:77–88. Sika, A. 1981. Present results of the
international provenance
of IUFRO with Douglas-fir in the CSR. Communicationes
Instituti Forestalis Cechosloveniae 12:83–101.
Sika, A. 1982. Dosavadni vysledky proveniencniho
vyzkumu doug- lasky tisoliste v CSR (with English
summary “Present results of the international
provenance experiment with Douglas-fir in the CSR).
Prace VULHM. Reports of the Forestry and Game
Management Research Institute Jiloviste–Strnady,
CSSR 60:7–25.
Sika, A. 1983. Douglas-fir production in the Czech
Socialist Republic. Communications Instituti Forestalis
Cechosloveniae 13:41–57.
Silen, RR. 1958. Artificial ripening of Douglas-fir cones.
Journal of Forestry 56:410–413.
Silen, RR. 1962a. Pollen dispersal considerations for Douglas-
fir. Journal of Forestry 60:790–795.
Silen, RR. 1962b. A study of the genetic control of bud
bursting in Douglas-fir. Journal of Forestry
60:472–475.
Silen, RR. 1965. Regeneration aspects of the 50-year-
old Doug- las-fir heredity study, pp. 35-39 in
Proceedings 1964 Annual Meeting, West. Refor.
Coord. Comm., Western Forest and Conservation
Association, Spokane, WA.
Silen, RR. 1966a. A simple progressive tree improvement
program for Douglas-fir. USDA Forest Service,
Pacific Northwest Forest and Range Experiment
Station Research Note PNW-45, Portland, OR. 13
p.
Silen, RR. 1966b. A 50–year racial study of Douglas-fir
in western Oregon and Washington, pp. 6–7 in
Proceedings 1965 Annual Meeting, Western Forest
Genetics Association.
Silen, RR. 1967. Earlier forecasting of Douglas-Fir cone
crop
using male buds. Journal of Forestry 65(12):888–892.
Silen, RR. 1973. First- and second-season effect on
Douglas- fir cone initiation from a single shade
period. Canadian Journal of Forest Research
3:528–534.
Silen, RR. 1978. Genetics of Douglas-fir. USDA Forest
Service, Research Paper WO-35. Washington, DC.
34 p.
Silen, RR. 1989. A clinal model of tree improvement proposed
for Douglas-fir. Unpublished manuscript. 39 p.
Silen, RR, and DL Copes. 1972. Douglas-fir seed
orchard prob- lems: A progress report. Journal of
Forestry 70(3):145–147.
Silen, RR, and G Keane. 1969. Cooling a Douglas-fir seed
orchard to avoid pollen contamination. Research
Note PNW 101, Pacific Northwest Forest and
Range Experiment Station, USDA Forest Service,
Corvallis, OR. 10 p.
Silen, RR, and JG Wheat. 1979. Progressive tree
336 Douglas-fir: The Genus Pseudotsuga
Silen, RR, DL Olson, and JC Weber. 1993. Genetic variation pression and growth promotion in Douglas-fir seedlings
in susceptibility to windthrow in young Douglas-fir.
Forest Ecology and Management 61(1):17–28.
Silim, SN, and DP Lavender. 1992. Relationships between
cold hardiness, stress resistance and bud dormancy in
white spruce (Picea glauca (Moench) Voss) seedlings,
pp. 9–14 in Proceedings of the 1991 Forest Nursery
Association of BC annual meeting, Prince George, BC
Canada, comp. FP Don- nelly and HW Lussenburg,
Forest Nursery Association of British Columbia.
Silver, GT. 1962. The distribution of Douglas-fir foliage by
age. Forestry Chronicle 38(4):433–438.
Simmard, SW, DA Perry, and JE Smith. 1997. Effects of soil
trenching on occurrence of ectomycorrhizas on Pseu-
dotsuga menziesii seedlings grown in mature forests of
Betula papyrifera and Pseudo menzeisii. New
Phytologist 136:327–340.
Simard, SW, DA Perry, MD Jones, DD Myrold, DM Dorrell,
and R Molina. 1997. Net transfer of carbon between
ecto- mycorrhizal tree species in the field. Nature
388(7):579–582.
Siminovitch, D, CM Wilson, and DR Briggs. 1953.
Studies on the chemistry of the living bark of the
black locust in relation to its frost hardiness. V.
Seasonal transforma- tions and variations in the
carbohydrates: starch-sucrose interconversions. Plant
Physiology 28:383–400.
Simonson, GH. 1963. Temperature and the water balance
for Oregon weather stations. Special Report 150.
Agriculture Experiment Station, Oregon State
University, Corvallis. 127 p.
Simpson, DG. 1986. Auxin stimulates lateral root growth of
interior Douglas-fir seedlings. Canadian Journal of
Forest Research 16:1135–1139.
Simpson, GM. 1990. Seed Dormancy in Grasses. Cambridge
University Press, Cambridge.
Şimşek, Y. 1977. Duglas (Pseudotsuga menziesii (Mirb.)
Franco in Türkiye’ye ithali ve orijin problemeri üzerine
arastirma- lar. (With German summary: Die Einführung
der Douglasie (Pseudotsuga menziesii (Mirb.) Franco in
die Türkei und deren Herkunftsfrage.) Poplar and Fast
Growing Forest Trees Research Institute, Izmit, Turkey,
Annual Bulletin No. 12, Yenilik Basimevi, Istanbul.
Şimşek, Y. 1978. Ergebnisse aus dem internationalen
Douglasienherkunftsversuch in der Türkei, pp. 357–368
in Vol. 1: Background papers and Douglas-fir provenances.
Pro- ceedings IUFRO Joint meeting of Working Parties
S2.02–05,
–08, –12, –14. Vancouver, B.C. 1978. B.C. Ministry of Forests,
Information Services Branch.
Şimşek, Y. 1980. Ergebnisse aus dem internationalen
Douglasien-Herkunftsversuch in der Türkei. Silvae
Ge- netica 29(2):44–50.
Şimşek, Y. 1982. 1973–1974 Yilinda Türkiye’de tesis edilen
ulus- lararasi duglas (Pseudotsuga menziesii (Mirb.)
Franco) orijin denemelerinin sonuclari. (With German
summary “Ergebnisse des internationalen Douglasien-
Provenienzversuches von 1973/74 in der Türkei. Poplar and
Fast Growing Exotic For- est Trees Research Institute,
Annual Bulletin No.17:1–36. Yenilik Basimevi, Istanbul.
Şimşek, Y. 1987. Karadeniz bölgesinde yapilacak duglas (Pseu-
dotsuga menziesii Mirb. Franco) agaclandirmalari
icinorijin secimi. (With German summary “Zur
Herkunftswahl bei Douglasienaufforstungen im
Schwarzmeergebiet der Türkei.”) Ormancilik Arastirma
Enstitüsü Yainlari, Teknik Bülten Serisi No. 190, Forest
Research Institute Ankara.
Sinclair, WA, DP Fowles, and SM Hee. 1975. Fusarium root
rot of Douglas-fir seedlings suppression by soil
fumigation, fertility management, and inoculation with
spores of the fungal symbiote Laccaria laccata. Forest
Science 21:390-399.
Sinclair, WA, DH Sylvia, and AO Larson. 1982. Disease sup-
by the ectomycorrhizal fungus Laccaria laccata. Soest, J. Van. 1954. Stormschade aan Douglas (with Engl.
Forest Science 28(2):191–201. Summary: Storm damage to Douglas-fir. Ned. Bosb.
Singh, P. 1970. Armillaria root rot in a coniferous Tijdschr. 26(4):89–99.
plantation
in Newfoundland. Bi-monthly Research Notes 26(1):
5–6. Singh, H. 1978. Embryology of Gymnosperms. 1978.
XII, Gebrüder
Borntraeger, Berlin. 302 p.
Sirakov, G. 1948. (The problem of the conversion of
beech forests on northern slopes of Stara Planina
(Balkan Moun- tains) into mixed forest with
conifers predominating). Sbornik na Centralnija
Gorski Izledovatelski Institut 1948: 141–206.
(Original not seen, cited from Forestry Abstracts
14:1098, 1953).
Sissons, HW. 1933. Distribution and rate of fall of conifer seeds.
Journal of Agricultural Research 44(2):119–128.
Skadsen, R. 1975. Synergisms between gibberellins
and gir- dling in promoting flowering in Douglas
fir grafts. Paper presented at Western Forest
Genetics Association Meeting, Duncan, August,
1975.
Skrzypczynska, M. 1994, A new viewpoint on the
development of Megastigmus spermotrophus
Wachtl, 1893 (Hymenop- tera, Torymidae) – a pest
in Douglas-fir seed. Wiadomosci Entomologiczne
13(3):181–183.
Skutko, NV. 1966. Varieties of Douglas-fir in forest
planta- tions in Belorussia. Rastitelnye Resursy
Izdataistvo “Nauka” 2(1):115–121.
Slanky, V. 1973. Hormonal relationships in Mycorrhizal
devel- opment, pp. 282–298 in Ectomycorrhizae Their
Ecology and Physiology. ed. GC Marks and TT
Kozlowski. Academic Press, NY. 444 p.
Smith, CC, JL Hemrick, and CL Kramer. 1990. The
advantage of most years for wind pollination. The
American Natural- ist 136(2):154–166.
Smith, H, and NP Kefford. 1964. The chemical
regulation of the dormancy phases of bud
development. American Journal of Botany
51:1002–1012.
Smith, JHG, and DL Reukema. 1986. Effects of
plantation and juvenile spacing on tree and stand
development, pp. 239–245 in Douglas-Fir: Stand
Management for the Future, ed. CD Oliver, DP
Hanley, and JA Johnson. Contribu- tion No. 55,
College of Forest Resources, University of
Washington, Seattle.
Smith, JHG, J Walters, and A Kozak. 1968. Influences of
Fertil- izers on Cone Production and Growth of Young
Douglas-fir, Western Hemlock, and Western Red Cedar
on the UBC Research Forest. University of British
Columbia, Faculty of Forestry Bulletin No. 5,
Vancouver, BC. 57 pages.
Smith R, and M Greenwood. 1995. Effects of gibberellin
A4/7, root pruning and cytokinins on seed and
pollen cone production in black spruce (Picea
mariana). Tree Physiol- ogy 15:457–465.
Smith, RF. 1998. Effects of stem injections of
gibberellin A4/7 and paclobutrazol on sex
expression and the within-crown distribution of seed
and pollen cones in black spruce (Picea mariana).
Canadian Journal of Forest Research 28:641–651.
Smith, WH, and RG Stanley. 1969. Cone distribution in
crowns of slash pine in relation to stem, crown and
wood incre- ments. Silvae Genetica 18:57–66.
Smucker, AJM, and RM Aiken. 1992. Dynamic root responses
to water deficits. Soil Science 154(4):280-289.
Soest, J Van. 1956. Niederländische Erfahrungen mit
aus- laendischen Nadelholzarten. Allgemeine
Forstzeitschrift 11(45/46):593–596.
Soest, J Van. 1959. Erfahrungen mit der Douglasie in
der nie- derländischen Forstwirtschaft. Allgemeine
Forstzeitschrift 14(8):155–157.
References 337
Spengler. 1925. Douglasienanbau im Büdinger Stadtwald.
Sokolowski, S. 1912. Hodowla lasu. Daglezya zielona, Da-
glezya szara. Original not seen, cited from Mejnartowicz
1976.
Soljanik, I. 1968. The frost resistance of Pseudotsuga
menziesii
var. viridis in Kosovo and Metohija. Sum. List 92(3/4):112–
121. Original not seen. Cited from Forestry Abstracts
30:3805, 1969.
Somerville, W. 1904. The Douglas-fir plantation at Taymount
(xxxii) abstracted from a paper entitled “Exotic conifers
in Britain.” Royal Scottish Arboricultural Society
Transac- tions 17(2):269–276.
Sorensen, FC. 1967. Two–Year Results of a West–East Transect
Provenance Test of Douglas-fir in Oregon. USDA, Forest
Service Research Note PNW–72, Portland, OR. 8 p.
Sorensen, FC. 1971. “White Seedling:” a pigment mutation
that affects seed dormancy in Douglas-fir. Journal of
He- redity 62(2):127–130.
Sorensen, FC. 1972. The Seed Orchard Tree as a Pollen
Sam- pler: A model and Example. USDA Forest
Service, Pa- cific Northwest Forest Experiment Station
Research Note 175:1–11.
Sorensen, FC. 1979. Provenance variation in Pseudotsuga
menziesii seedlings from the var. menziesii–var. glauca
transition zone in Oregon. Silvae Genetica 28(2/3):96–
103.
Sorensen, FC. 1980. Effect of Date of Cone Collection and
Strati- fication Period on Germination and Growth of
Douglas-fir Seeds and Seedlings. USDA Forest Service
Research Note PNW 346, Pacific Northwest Forest and
Range Experi- ment Station, Portland, OR. 11 p.
Sorensen, FC. 1983. Geographic variation in seedling
Douglas- fir (Pseudotsuga menziesii) from the Western
Siskiyou Mountains in Oregon. Ecology 64(4):696–702.
Sorensen, FC. 1991. Stratification Period and Germination of
Douglas-fir Seed From Oregon Seed Orchards, Two Case
Studies. USDA Forest Service Pacific Northwest Research
Station Research Note PNW–RN–499, Portland, OR. 22 p.
Sorensen, FC. 1996. Effects of length of seed chilling
period and sowing date on family performance and
genetic variances of Douglas-fir seedlings in the
nursery. New Forests 12(3):187–202.
Sorensen, FC. 1999. Effect of dry storage on germination rate of
seeds of coastal Douglas-fir (Pseudotsuga menziesii)
(Mirb) Franco var. menziesii. Seed Science and Technology
27, 91–99.
Sorensen, FC, and RK Campbell. 1981. Germination rate of
Douglas-fir (Pseudotsuga menziesii) (Mirb.) Franco
seeds affected by their orientation. Annals of Botany
44:467–471.
Sorensen, FC, and RK Campbell. 1985. Effect of seed weight
on height growth of Douglas-fir (Pseudotsuga menziesii
(Mirb.) Franco var. menziesii) seedlings in a nursery.
Canadian Journal of Forest Research 15:1109–15.
Sorensen, FC, and RK Campbell. 1993. Seed weight-seedling
size correlation in coastal Douglas-fir: genetic and en-
vironmental improvements. Canadian Journal of Forest
Research 23(2):275–285.
Sorensen, FC, and RS Miles. 1978. Cone and seed weight
relationships in Douglas-fir from western and central
Oregon. Ecology 59:641–644.
Sornay, J. 1937. Les peuplements de Douglas en France.
Bulletin de la Société Forestière de Franche-Comté Salin
les Bains 22(4):198–214.
Soutrenon, A. 1986. Observations sur le comportement de
provenance de douglas vert et de douglas bleu à une
attaque de Rhabdocline. Forêts de France No.291a:10–
16.
Spencer, DA, and NB Kverno. 1953. Research in rodent
control to promote reforestation by direct seeding.
Progress Report No. 4, Sec. 2, USDI Fish and Wildlife
Service, Denver Wildlife Research Laboratory, 7 p.
Stein, JD, and GP Markin. 1986. Evaluation of four
Allgemeine Forst– und Jagdzeitung 101:420–421. chemi- cal insecticides registered for control of the
Douglas-fir cone gall midge, Contarinia oregonensis
Sprague, J, JB Jett, and B Zobel. 1979. The mangement (Diptera: Ceci- domyiidae), and the Douglas-fir seed
of southern pine seed orchards to increase seed chalcid, Mega- stigmus spermotrophus (Hymenoptera:
production, pp. 145–162 in Proceedings Symposium Torymidae), in Douglas-fir seed orchards. The
Flowering and Seed Development, Mississippi State Canadian Entomologist 118(11):1185–1191.
University, 1978. ed. F Bon- ner, IUFRO and USDA
Forest Service Southern Forest Experiment Station,
Starkville, MS.
Spurr, SH. 1961. Observations on Douglas-fir in New
Zealand. Forest Research Institute, Technical Paper
No. 38, New Zealand Forest Service, Rotorua.
Squillace, AE, and Long, EM. 1981. Proportion of
pollen from non-orchard sources, pp. 15–19 in
Pollen Management Handbook, ed. EC Franklin.
USDA Agriculture Handbook
587. Washington, DC.
St. Clair, J.B. 1994. Genetic variation in tree structure
and its relation to size in Douglas-fir. II. Crown
form branch characters, foliage characters.
Canadian Journal of Forest Research 24:1236–
1247.
St. Clair, JB, NL Mandel, and KJS Jayawickrama. 2004.
Early realized genetic gains for coastal Douglas-fir
in the north- ern Oregon Cascades West. Western
Journal of Applied Forestry 19(3) :195–201.
St. Clair, JB, NL Mandel, KW Vance-Borland. 2005.
Genecol- ogy of Douglas-fir in western Oregon and
Washington. Annals of Botany 96(7):1199–1214.
St. Clair, JB, and WT Adams. 1991. Effects of seed
weight and rate of emergence on early growth of
open-pollinated Douglas-fir families. Forest
Science 37:987–997.
St. John, TY, and DC Coleman. 1983. The Role of Mycorrhizae
in Plant Ecology. Canadian Journal of Botany
61:1005–1014. Stack, RW, and WA Sinclair. 1975.
Protection of Douglas-fir
seedlings against Fusarium root rot by a mycorrhizal
fungus in the absence of mycorrhiza formation.
Phyto- pathology 65:468–472.
Stadnitskii, GW. 1986. Cone and seed insect diapause
and its significance in seed damage. pp. 241–44 in
Proceedings of the 2nd Conference of the Cone and
Seed Insects Working Party S2.07–01. Briançon,
05100, France, September 3–5, 1986. Station de
Zoologie Forestière, INRA-CRF, Olivet, France.
Staebler, GR. 1956. Evidence of shock following
thinning of young Douglas-fir. Journal of Forestry
54(5):339.
Stamenkovic, V, and V Miscevic. 1973. Entwicklung
und Zuwachs einiger Nadelholzkulturen in
Zentral–und Ostserbien (with Engl. summary:
Development and incre- ment of some conifer
plantations in central and eastern Serbia).
Forstwissenschaftliches Centralblatt 9(5):257–291.
Standley, PC. 1920–26. Trees and shrubs of Mexico.
Contribu- tions from the U.S. National Herbarium
23:1–1721.
Stanley, RG, and EG Kirby. 1973. Shedding of pollen
and seeds, pp. 295–340 in Shedding of plant parts.
Academic Press, New York.
Stanley, RG, and HF Linskens. 1974. Pollen: biology,
biochem- istry, management. Springer-Verlag, New
York.
Stark, R. 1965. Recent trends in forest entomology.
Annual Review of Entomology l0:303–324.
Stein JD, and TW Koerber. 1989. Timing of acephate
implants for control of the coastal seed insect
complex of Douglas- fir in northern California, pp.
203–209 in Proceedings, 3rd Cone and Seed Insects
Working Party Conference, 1988 June 26–30,
Victoria, BC. USDA Forest Service, Pacific South-
west Forest and Range Experiment Station,
Berkeley, CA.
338 Douglas-fir: The Genus Pseudotsuga
Stein JD, TW Koerber, and CL Frank. 1988. Trunk- netic gain estimates in coastal Douglas-fir in British Columbia.
implanted Acephate to protect Douglas-fir seed crops on British Columbia Ministry of Forests and Range, Forest
individual trees in northern California. Journal of
Economic Entomol- ogy 81(6):1668–1671.
Stein, JD, RE Sandquist, TW Koerber, and CL Frank. 1993.
Response of Douglas-fir cone and seed insects to
implants of systemic insecticides in a northern California
forest and a southern Oregon seed orchard. Journal of
Economic Entomology 86(2):465–469.
Stein, WI. 1963. Plecoma larvae, root feeders in Western
forests.
Northwest Science 37(4):126–143.
Stein, WI. 1965. Field test of Douglas-fir, ponderosa pine,
and sugar pine seeds treated with hydrogen peroxide.
Tree Planters’ Notes 71:25–29.
Stein, WI. 1966. Sampling and Service Testing Western
Conifer Seeds. 86 p. Western Forest Tree Seed
Council/Western Forestry and Conservation Committee,
Portland, OR. 36 p.
Stein, WI. 1967. Laboratory seed tests–are they doing the
job? pp. 20–23 in Proceedings of the 1967 meeting of
the Western Reforestation Coordinating Committee.
Western Forestry and Conservation Association.
Portland, OR.
Stein, WI, and PW Owston. 2008. Pseudotsuga Carr. Douglas-
fir, pp. 891-906 in The Woody Plant Seed Manual: A
handbook on seeds of trees and shrubs, ed. FT Bonner
and RP Kar- rfalt. USDA Forest Service Agriculture
Handbook 727, Washington, DC.
Stein, WI, Danielson, R, N Shaw, S Wolff, and D Gerdes.
1986. Users Guide for Seeds of Western Trees and
Shrubs. USDA Forest Service, General Technical Report
PNW-193, Pacific Northwest Research Station, Portland,
OR. 45 p.
Steinbrenner, EC, and SP Gessel. 1956. Windthrow along
cut- lines in relation to physiography on the McDonald
tree farm. Weyerhaeuser Timber Company Forest
Research Note 15. 19 p.
Steinbrenner, EC, and JH Rediske. 1964. Growth of
ponderosa pine and Douglas-fir in a controlled
environment. Weyer- haeuser Forest Paper No. 1,
Centralia, WA. 31 p.
Steinbrenner, EC, JW Duffield, and RK Campbell. 1960. In-
creased cone production of young Douglas-fir following
nitrogen and phosphorus fertilization. Journal of
Forestry 58:105–110.
Steiner, KC, and JW Wright. 1975. Douglas-fir Christmas
trees: Variation in frost susceptibility and time of
leafing out in Michigan. Michigan Academy 7(2):185–
190.
Steinfeld, D, MP Amaranthus, and E Cazares. 2003.
Survival of ponderosa pine (Pinus ponderosa Dougl.
ex Laws.) seedlings outplanted with Rhizopogon
mycorrhizae inocu- lated with spores at the nursery.
Journal of Arboriculture 29(4):197–208.
Stephan, BR. 1987. Differences in the resistance of Douglas-
fir provenances to the woolly aphid Gilletteella cooleyi.
Silvae Genetica 36(2):76–79.
Sterling, C. 1963. Structure of the male gametophyte in gym-
nosperms. Biological Reviews 38(2):167–203.
Stern, K, A König, and HH Hattemer. 1974. Beitraege
zum geographischen Variationsmuster der Douglasie.
Silvae Genetica 23(1/3):53–58.
Stimm, B, and PH Dong. 2001. Der Douglasien–
Herkunftsver- such Kaiserslautern nach neun
Jahrzehnten Beobachtung (with Engl. summary: The
Kaiserslautern Douglas-fir provenance trial after nine
decades of observation). Forst- wissenschaftliche
Centralblatt 120:173–186.
Stoate, TN, I Mahood, and EC Croesin. 1961. Cone produc-
tion in Douglas-fir (Pseudotsuga menziesii). Empire
Forestry Review 40:104–110.
Stoehr, M, J Woods, K Bird, and L Hayton. 2011. Verifying ge-
Science Progress Extension Note 104, Victoria, BC. in Grosspolen). Roczniki Nauk Rolniczych i Lesnych
5 p. Stoehr, MK, K Bird, G Nigh, J Woods, and A (Annals of the Agriculture and Forestry Sciences) Vol.
15.
Yanchuk. 2010.
Realized genetic gains in coastal Douglas-fir in British Sud’ev, NG. 1980. (Fast–growing species in plantations in the
Columbia: implications for growth and yield projections.
Silvae Genetica 59(5):223–233.
Stoleson, RM, and RG Hallman. 1972. Helicopter seeders can
damage conifer seed. Tree Planters’ Notes 23(3) 28.
Stone, EC. 1967. A nursery-conditioned root growth
response to the field environment, pp. 282–286 in
Proceedings XIV International Union of Forestry
Research Organizations (IU- FRO) Congress,
Munich.
Stone, EC, and GH Schubert. 1958. Seasonal periodicity
in root regeneration of ponderosa pine transplants—
a physi- ological condition. Proceedings, Society of
American Foresters 1958, pp. 154–155.
Stone, EC, and GH Schubert. 1959a. Root regeneration
by ponderosa pine seedlings lifted at different times
of the year. Forest Science 5(4):322–332.
Stone, EC, and GH Schubert. 1959b. The physiological
condi- tion of ponderosa pine (Pinus ponderosa
Laws.) planting stock as it affects survival after cold
storage. Journal of Forestry 57:837–841.
Stone, EC, and GH Schubert. 1959c. Root
regeneration by seedlings: Ability of ponderosa
pine seedling to regener- ate root system rapidly
after transplanting is important factor in survival.
California Agriculture 13(2):12,14.
Stone, EC, JL Jenkinson, and SL Krugman. 1962. Root
Regener- ating potential of Douglas-fir seedlings
lifted at different times of the year. Forest Science
8(3):288–297.
Stonecypher, RW, RF Piesch, GG Helland, JG
Chapman, and HJ Reno. 1996. Results from genetic
tests of selected par- ents of Douglas-fir
(Pseudotsuga menziesii [Mirb.] Franco) in an
applied tree improvement program. Forest Science
42(2): Monograph 32. 21 p.
Strasburger, E. 1872. Die Coniferen und die Gnetaceen.
Jena. Strasburger, E. 1887. Das Botanische Practicum.
Jena.
Strauss, SH, AH Doerksen, and JR Byrne. 1990.
Evolution- ary relationships of Douglas-fir and its
relatives (genus Pseudotsuga) from the DNA
restriction fragment analysis. Canadian Journal of
Botany 68:1502–1510.
Streets, RJ. 1962. Exotic forest trees in the British Commonwealth.
Clarendon Press, Oxford.
Stringer, PW, and GH LaRoi. 1970. The Douglas-fir
forests of Banff and Jasper National Parks, Canada.
Canadian Journal of Botany 48:1703–1726.
Strobel, NE, and WA Sinclair. 1991a. Role of
flavonolic wall infusions in the resistance induced
by Laccaria bicolor to Fusarium oxysporum in
primary roots of Douglas-fir. Phytopathology
81:420-425.
Strobel, NE, and WA Sinclair. 1991b. Influence of
tempera- ture and pathogen aggressiveness on
biological control of Fusarium root rot by Laccaria
bicolor in Douglas-fir. Phytopathology 81(4):415–
420.
Studholme, WP. 1995. The experience and
management strategy adopted by the Selwyn
Plantation Board, New Zealand, pp. 468–476 in
Wind and Trees, ed. MP Coutts and J Grace,
Cambridge University Press, Cambridge.
Sturrock, R., and R. Garbutt. 1994. Laminated root rot of
Douglas- fir. Forest Pest Leaflet, Canadian Forest
Service, Pacific Forestry Centre, Natural Resources
Canada, Victoria, BC. 8 p.
Suchocki, S. 1926. Pseudotsuga Douglasii i
dotychczasowe wyniki jej alkimatyzacji w
Poznanskiem (with German summary: Pseudotsuga
douglasii und die Ergebnisse ihrer Anbauversuche
References 339
Zealand Forest Service, Rotorua.
Carpathians and the Caucasus). Lesnaya Sweet, GB. 1965. Provenance Differences in Pacific Coast
Promyshlennost’ 1980 No. 7:9–10. Original not seen, Douglas-fir. Silvae Genetica. 14(2):46–56.
cited from Forestry Abstracts 42:1007, 1981.
Sudworth, GB. 1908. Forest Trees of the Pacific Slope, USDA,
Forest Service Washington, DC, pp. 104–106.
Sudworth, GB. 1918. Miscellaneous Conifers of the Rocky
Moun- tain Region. USDA Forest Service Bulletin 680,
Washing- ton, DC.
Summers, D, and GE Miller. 1986. Experience with systemic
insecticides for control of cone and seed insects in Doug-
las-fir seed orchards in coastal British Columbia,
Canada. pp. 267–283 in Proceedings of the 2nd
Conference of the Cone and Seed Insects Working Party
S2.07-01 Briancon, 05100, France, September 3-5,
1986, ed. A Roques. International Union of Forestry
Research Organizations, Olivet, France.
Summers, D, and DS Ruth. 1987. Effect of diatomaceous
earth, Malathion, Dimethoate and Permethrin on
Leptoglossus occidentalis (Hemiptera: Coreidae): A pest
of conifer seed. The Journal of the Entomological Society
of British Columbia 84:33–38.
Susmel, L. 1962. La Douglasia verde. Monti e Boschi
13(11/12):579–590.
Sutherland, JR. 1979. The pathogenic fungus Caloscypha
fulgens in stored conifer seeds in British Columbia and
relation of its incidence to ground and squirrel-cache
collected cones. Canadian Journal of Forest Research
9:129–132.
Sutherland, JR, and TAD Woods. 1978. The fungus
Geniculo- dendron pyriforme in stored Sitka spruce
seeds: Effects of seed extraction and cone collection
methods on disease incidence. Phytopathology 68:747–
750.
Sutherland, JR, and E VanEerden. 1980. Diseases and Insect
Pests in British Columbia Forest Nurseries. British
Colum- bia Ministry of Forests, Canadian Forestry
Service, Joint Report 12, Victoria, BC. 55 p.
Sutherland, JR, W Lock, and SH Farris. 1981. Sirococcus
blight: a seed-borne disease of container-grown spruce
seedlings in Coastal British Columbia forest nurseries.
Canadian Journal of Botany 59:559–562.
Sutherland, JR, T Miller, and RS Quinard. 1987. Cone and
Seed Diseases of North American Conifers. North
American Forestry Commission Publication Number 1.
Victoria, BC. 77 p.
Sutton, RF. 1979. Planting stock quality and grading. Forest
Ecology and Management 2:123–132.
Sutton, RF. 1990. Root growth capacity in coniferous
forest
trees. Horticultural Science 25(3):259–266.
Sutton, RF. 1991. Soil Properties and Root Development in
Forest Trees: A Review. Great Lakes Forestry Centre,
Information Report O-X-413, Ontario Region, Forestry
Canada. 42 p.
Sutton, RF. 1994. Field Stop CP68. Roots and Shoots: A
Reforesta- tion Workshop. Canadian Forest Service, Prince
George District Office, British Columbia.
Sweeney, JD, and GE Miller. 1989. Distribution of Barbara
col- faxiana (Kearfott) (Lepidoptera: Tortricidae) eggs
within and among Douglas-fir crowns and methods for
estimat- ing egg densities. The Canadian Entomologist
121:569–578.
Sweeney, JD, YA El-Kassaby, DW Taylor, DGW Edwards,
and GE. Miller. 1991. Applying the IDS method to
remove seeds infested with the seed chalcid,
Megastigmus sper- motrophus in Douglas-fir,
Pseudotsuga menziesii (Mirb.) Franco. New Forests
5:327–334.
Sweet, GB. 1964. The Establishment of Provenance Trials in
Douglas-fir (Pseudotusga menziesii (Mirb.) Franco) in
New Zealand. Forest Research Institute, Genetics and
Tree Improvement Report No. 21 (unpublished), New
Takaso T, P von Aderkas, and JN Owens. 1996. Prefertiliza-
Sweet, GB. 1975. Flowering and seed production, pp. tion events in ovules of Pseudotsuga: ovular secretion
72–82 in Seed Orchards, ed. R. Faulkner, UK and its influence on pollen tubes. Canadian Journal of
Forestry Commission Bulletin 54, HMSD, London. Botany 74:1214–1219.
Sweet, GB, and MP Bollman. 1972. Regional variation Tanaka, Y. 1984. Assuring seed quality for seedling
in Douglas-fir seed yields. New Zealand Journal of production: Cone collection and seed processing, testing,
Forestry 17(1):74–80 storage and
Sweet, GB, and GM Will. 1965. Precocious male cone
produc- tion associated with low nutrient status in
clones of Pinus radiata. Nature (London) 206:739.
Sylvia, DM. 1983. Role of Laccaria laccata in
protecting pri- mary roots of Douglas-fir from root
rot. Plant and Soil 71:299–302.
Sylvia, DM, and WA Sinclair. 1983a. Suppressive
influence of Laccaria laccata on Fusarium oxysporum
and on Douglas-fir seedlings. Phytopathology 73:384–
389.
Sylvia, DM, and WA Sinclair. 1983b. Phenolic
compounds and resistance to fungal pathogens
induced in primary roots of Douglas-fir seedlings
by the ectomycorrhizal fungus Laccaria laccata.
Phytopathology 73:390-397.
Sziklai, O. 1969. Preliminary Notes on Variations in
(some) Cone and Seed Morphology of Douglas-fir.
Second World Con- ference on Forest Tree Breeding,
Washington, DC. FO– FTB–69–6/9:11.
Sziklai, O. 1990. Douglas-fir provenance/progeny test at
Haney, Brit. Columbia, Canada. Paper 2.313 in
Proceedings Joint Meeting WFGA and IUFRO
Working Parties S2.02–05, 06, 12, 14, Olympia,
Washington. Weyerhaeuser Company, Centralia,
WA.
Sziklai, O, YA El–Kassaby, and RK Scagel. 1987.
Relationship of Pseudotsuga menziesii with Other
Pseudotsuga Species Inferred From Karyotype
Reconstruction, pp. 127–142 in FBVA–Berichte
No. 21, Schriftenreihe der Forstlichen
Bundesversuchsanstalt, Vienna, Austria.
Szönyi, L. 1963. A hazai duglásfenyö–állományok
termöhelyi viszonyai (with German summary: Die
Standortverhaelt- nisse der einheimischen
Douglasienbestaende). Magyar Tudományos
Akademia Agrartudomanyok Osztalyanak Közle- menyei
22(1/2):79–92.
Szönyi, L, and I Nagy. 1968. Frostresistenz und
Wachstum der Douglasie, pp. 68–76 in
Klimaresistenz, Photosynthese und
Stoffproduktion. Tagungsbericht No. 100
Deutsche Akademie der
Landwirtschaftswissenschaften, Berlin, Deutsche
Demokratische Republik.
Tabbush, PM. 1986. Rough handling, soil temperature,
and root development in outplanted Sitka spruce and
Douglas fir. Canadian Journal of Forest Research
16:1385–1386.
Tabbush, PM. 1987. Effect of desiccation on water
status and forest performance of bare rooted Sitka
spruce and Douglas-fir transplants. Forestry
60(1):31–43.
Taimre, H. 1954. Ebatsuuga (douglaasia) kasvatamise
voim- alusi Eesti NSV metsades. Eesti NSV
Teaduste Akadeemia Toimetised, III kd, No. 4
(original not seen, cited from Margus 1961).
Takaso, T, and JN Owens. 1994. Effects of ovular
secretions on pollen in Pseudotsuga menziesii
(Pinaceae). American Journal of Botany 81:504–
513.
Takaso, T, and JN Owens. 1995. Ovulate cone
morphology and pollination in Pseudotsuga and
Cedrus. International Journal of Plant Sciences
156:630–635.
Takaso, T, and JN Owens. 1996. Postpollination-
prezygotic ovular secretions into the micropylar
canal in Pseudotsuga menziesii (Pinaceae). Journal
of Plant Research 109:147–160.
340 Douglas-fir: The Genus Pseudotsuga
stratification, pp. 27–40 in Forest Nursery Manual: Pathology 34(6):383–394.
Produc- tion of Bare Root Seedlings, ed. ML Duryea and
TD Landis, Forest Research Laboratory, OSU, Corvallis,
OR/ Martinus Nijhoff/Dr. W. Junk Publishers, Boston,
MA. 385 p.
Tappeiner, JC. 1969. Effect of cone production on branch,
needle, and xylem ring growth of Sierra Nevada
Douglas- fir. Forest Science 15:171–174.
Tappeiner, JC, D Huffman, D Marshall, TA Spies, and JD
Bai- ley. 1997. Density ages, and growth rates in old-
growth and young growth forests in coastal Oregon.
Canadian Journal of Forest Research 27: 638–648.
Tappeiner, JE III. 1969. Effect of cone production on branch,
needle, and xylem ring growth of Sierra Nevada
Douglas- fir. Forest Science 15(2):171–174.
Tarrant, RF. 1953. Effect of Heat on Soil Color and pH of
Two Forest Soils. Research Note PNW-90. USDA Forest
Ser- vice, Pacific Northwest Range and Experiment
Station, Portland, OR.
Tarrant, RF. 1954. Soil Reaction and Germination of
Douglas-fir Seed, Research Note 105, USDA Forest
Service, Pacific Northwest Range and Experiment
Station, Portland, OR.
Tarrant, RF. 1956. Effects of slash burning on some soils of
the Douglas-fir region. Proceedings Soil Science Society
of America 20:408–411.
Tauscher, AM. 1806. Lepidopterorom Russie Indigeorum
Ob- servatsiones sez. Mem.Soc. Nat. Moscowi:(174–
179):207–212
Tavoda, P. 1991. Zhodnotenie 20–rocneho proveniencneho
pokusu s duglaskou na slovensku (with German sum-
mary: “Auswertung eines zwanzigjaehrigen Provenien-
zversuches mit der Douglasie in der Slowakei”).
Vedecke Prace Vyskumneho Ustavu Lesneho Hospodarstva
Vo Zvolene 1991:131–146.
Taylor, AH, and CN Skinner. 1998. Fire history and
landscape dynamics in a late successional reserve,
Klamath Moun- tains, California, USA. Forest Ecology
and Management 111:285–301.
Taylor, JS, and PF Wareing. 1979a. The effect of
stratification on the endogenous levels of gibberellins
and cytokinins in seeds of Douglas-fir (Pseudotsuga
menziesii (Mirb.) Franco) and sugar pine (Pinus
lambertiana Dougl.) Plant, Cell & Environment
2(2):165–171.
Taylor, JS, and PF Wareing. 1979b. The effect of light on the
endogenous levels of cytokinins and gibberellins in
seeds of sitka spruce (Picea sitchensis Carriere). Plant,
Cell & Environment 2(2):173–179.
Taylor, MT, R Rose, DL Haase, and ML Cherry. 2011.
Effects of plant date and nursery dormancy induction on
field performance of Douglas-fir seedlings in western
Oregon. Tree Planters’ Notes 54(2/Fall):50–64.
Taylor, MA, HV Davies, SB Smith, A Abruzzese, and PG
Gosling. 1993. Cold-induced changes in gene expres-
sion during dormancy breakage in seeds of Douglas-
fir (Pseudotsuga menziesii). Journal of Plant
Physiology 142(1):120–123.
Taylorson, RB, and SB Hendricks. 1977. Dormancy in seeds.
Annual Review of Plant Physiology 28:331–354.
Telewski, FW, and MJ Jaffe. 1986. Thigmomorphogenesis:
Anatomical, morphological and mechanical analysis of
genetically different sibs of Pinus taeda in response to
me- chanical perturbation. Physiologia Plantarum
66(2):219–226.
Telewski, FW. 1990. Growth, wood density, and ethylene
pro- duction in response to mechanical perturbation in
Pinus taeda. Canadian Journal of Forest Research
20:1277–1282.
Temel, F, GR Johnson, and JK Stone. 2004. The relationship
between Swiss needle cast symptom severity and level of
Phaeocryptopus gaeumannii colonization in coastal
Douglas- fir (Pseudotsuga menziesii var. menziesii) Forest
Temel, F, GR Johnson, and WT Adams. 2005. Early Washington, Seattle.
genetic testing of coastal Douglas-fir for Swiss Thulin, I. 1949. Beskadigelser af douglasgran (Pseudotsu-
needle cast toler- ance. Canadian Journal of Forest ga taxifolia) i Danmark i vinteren 1946-47 (with Engl.
Research 35(3): 521–529. summary: Damage to Douglas fir in Denmark in the
Theisen, PA. 1976. Damage caused by the western seed
bug - Leptoglossus occidentalis. USDA Forest
Service Tree Seed Training Packet. Handout No. 6.
Pacific Northwest Re- gion, USDA Forest Service,
Portland, OR. 4 p.
Theisen, PA. 1980. Artificial Ripening of Conifer Seeds.
USDA
Forest Service Tree Seed Training Packet. Handout
No.
14. Pacific Northwest Region, USDA Forest
Service, Portland, OR. 9 p.
Theisen, PA, and D Goheen. 1980. Cone and Seed Disease. USDA
Forest Service Region 8. Tree Seed Training Packet. 10 p.
Thies, WG. 1983. Determination of growth reduction in
Douglas-fir infected by Phellinus weirii. Forest
Science 29(2):305–315.
Thies, WG. 1995. Species manipulation as a strategy to
reduce the impacts of lammeted root rot in
regenerated coastal strands. Cope Report 8(4):7–9.
Thies, WG, and EM Goheen. 2002. Major Forest
Diseases of the Oregon Coast Range and their
Management, pp. 191–210 in Forest and Stream
Management in the Oregon Coast Range, ed. SD
Hobbs, JP Hayes , RL Johnson, GH Reeves, TA
Spies, JC Tappeiner, GE Wells. Oregon State
University Press, Corvallis.
Thies, WG, and RN Sturrock. 1995. Laminated root rot in
Western North America. USDA Forest Service,
General Technical Report PNW-GTR 349, PNW
Research Station, Portland, OR. 32 pp.
Thies, WG, and DJ Westlind. 2005. Stump removal and
fertil- ization of five Phellinus weirii-infested
stands in Washing- ton and Oregon affect mortality
and growth of planted Douglas-fir 25 years after
treatment. Forest Ecology and Management
219(2/3):242–258.
Thies, WG, EE Nelson, and D Zobowski. 1994.
Removal of stumps from a Phellinus weirii infested
site and fertiliza- tion affect mortality and growth of
planted Douglas-fir. Canadian Journal of Forest
Research 24(2):234–239.
Thies, WG, KW Russell, and LC Weir. 1979. Rhizina
root rot of little consequence in Washington and
Oregon. Journal of Forestry 77(1):22–24.
Thomas, SC, and WE Winner. 2002. Photosynthetic
differ- ences between saplings and adult trees: an
integration of field results by meta-analysis. Tree
Physiology 22:117–127.
Thompson, BE, and RT Timmis. 1978. Root
regeneration potential in Douglas-fir seedlings:
effect of photoperiod and air temperature on its
evaluation and control, pp. 86–109 in Proceedings
IUFRO Symposium on Root Physiol- ogy and
Symbiosis, Nancy, France, September 11–15, 1978,
ed. A Riedacker and J. Gagnaire-Michard.
Thomson, AJ, and YA EL-Kassaby. 1993. Interpretation of
seed-germination parameters. New Forests 7(2):123–132.
Thomson, AJ, DG Goodenough, HJ Banllay, YJ Lee,
and RN Sturrock. 1996. Effects of laminated root
rot Phellinus weirii on Douglas-fir foliar chemistry.
Canadian Journal of Forest Research 26(8) 1440-
1445.
Thorbjornsen, E. 1856. Viability Tests and Pre-treatment of
Pon- derosa Pine Seeds from Northwest Coast Region.
MS thesis, University of Washington, Seattle.
Thornley, JHM. 1972. A model to describe the
partitioning of photosynthate during vegetative
plant growth. Annals of Botany 36:419–430.
Thrupp, AC. 1939. Effects of Seed Ash Characteristics and
Treat- ment of Seed and Soils upon Coniferous Seed
Germination. Doctoral dissertation, University of
References 341
character- ization of a set of cDNAs differentially
winter of 1946/47). Det forstlige Forsøgsvaesen i Danmark expressed during Douglas-fir germination and early
19(4):285–329. seedling development. Physiologia Plantarum 95:456–
464.
Thulin, IJ. 1967. Provenance testing, Pseudotsuga menziesii.
New Zealand Forest Service, Report of Forest Research Trappe, JM. 1961. Strong hydrogen peroxide for sterilizing coats
Institute for the period 1 January to 31 December 1966. of tree seed and stimulating germination. Journal of Forestry
59(11):828–829.
Tigerstedt, PM. 1990. Adaptability of seed sources across
geographic zones—90 years of experimenting in
Finland. Section 2.322 in Proceedings Joint Meeting of
Western Forest Genetics Association and IUFRO Working
Parties S2.02–05, 06, 12 and 14. Olympia, Washington,
Aug. 20–24, 1990.
Tillotson, CR. 1921. Storage of coniferous tree seed. Journal of
Agricultural Research XXII (9):479–510
Timmis, R. 1976. Methods of screening tree seedlings for
frost hardiness, pp. 421–435 in Tree Physiology and
Yield Improvement, ed. MGR Canell and FT Last.
Academic Press, London.
Timmis, R. 1974. Effect of nutrient stress on growth, bud set,
and hardiness in Douglas-fir seedlings, pp. 187-193 in
Proceedings of the North American Containerized forest
Tree Seedling Symposium, 26-29 Aug. 1974, Denver,
CO. Pub- lication 68, Great Plains Agricultural Council,
Denver.
Timmis, R. 1977. Critical frost temperatures for Douglas-fir
cone buds. Canadian Journal of Forest Research 7(1):19–
22.
Timmis, R, and Y Tanaka. 1976. Effects of container density
and plant water stress on growth and cold hardiness of
Douglas fir seedlings. Forest Science 22(2):167–172.
Timmis, R, and J Worrall. 1974. Translocation of
dehardening and bud break promoters in climatically
"split" Douglas- fir. Canadian Journal of Forest
Research 4(2):229–237.
Timmis, R, and J Worrall. 1975. Environmental control of
cold acclimation in Douglas-fir during germination, ac-
tive growth, and rest. Canadian Journal of Forest
Research 5: 464–77.
Timmis, R, J Flewelling, and C Talbert. 1994. Frost injury
prediction model for Douglas-fir seedlings in the Pacific
Northwest. Tree Physiology 14 (7/9):855–869.
Tinus, RW. 1992. Physiological testing of plants as a man-
agement tool, Combined Proceedings, International
Plant Propagators’ Society 42: 225–230.
Titze. 1906. Wachstumsleistungen von Pseudotsuga Douglasii
im
Sachsenwald. Zeitschrift für Forst-und Jagdwesen 38(8):536–
7. Titze. 1920. Wachstumsleistungen von Pseudotsuga Douglasii
im
Sachsenwald. Mitteilungen der Deutschen
Dendrologischen
Gesellschaft 29:269–270.
Tkatchenko, ME. 1958. Uldine metsakasvatus (General
Silvicul- ture. Translated from Obshee Lesovodsdvo,
Moscow–Leningrad 1955). Eesti Riiklik Kirjastus, Tallin,
Estonia.
Tompsett, PB, and AM Fletcher. 1979. Promotion of
flowering on mature Picea sitchensis by gibberellin and
environmen- tal treatments. The influence of timing and
hormonal concentration. Physiologia Plantarum
45(1):112–116.
Toumey, JW, and CL Stevens. 1928. The Testing of Conifer-
ous Tree Seeds at the School of Forestry. Yale
University, 1906–1926. Yale University School of
Forestry Bulletin No. 21, New Haven, CT. 46 p.
Toval, G. 1987. Results in two test sites of provenances of
the IUFRO collection of Pseudotsuga menziesii (Mirb.)
Franco in Galicia, pp. 49–66 in Proceedings of IUFRO
meeting, Breeding strategy for Douglas fir as an
introduced species, 1985 Vienna, Austria. FBVA–Berichte
No.21, Schriftenreihe der Forstlichen
Bundesversuchsanstalt, Vienna, Austria.
Tranbarger, TJ, and S Misra. 1995. The molecular
USDA Forest Service. 1975. Forest nursery diseases in the
Trappe, JM. 1977. Selection of fungi for United States. GW Peterson and RS Smith, Jr, tech.
ectomycorrhizal in- oculation in nurseries. Annual coords. USDA Forest Service, Agriculture Handbook
Review of Phytopathology 15:203–222. 470. Washington, DC. 125 p.
Trappe, JM. 1988. Lessons from alpine fungi. Mycologia 80:1-10.
Trappe, JM. 1989. The meaning of mycorrhizae to plant
ecol- ogy, pp. 347-349 in Mycorrhizae for Green
Asia: Proceedings of the First Asian Conference on
Mycorrhizae, January 29-31, 1988, Madras, India,
ed. M. Mahadevan, N. Ramen, and
K. Natarayan, Centre for Advanced Studies in Botany,
University of Madras. 351 p.
Trappe, JM, and RF Strand. 1969. Mycorrhizal
deficiency in a Douglas-fir region nursery. Forest
Science 15:381–389.
Treshow, M, FK Anderson, and F Harner. 1967.
Responses of Douglas-fir to elevated atmospheric
fluorides. Forest Science 13(2):114–120.
Trewin, HRD. 1978. What’s new in forest research,
pine seedlings-handle with care. Number 67,
Forest Research Institute, Rotorua.
Troup, RS. 1932. Exotic Forest Trees in the British Empire. Clar-
endon Press, Oxford. 259 p.
Tsukada, M. 1982. Pseudotsuga menziesii (Mirb.)
Franco): its pollen dispersal and late Quaternary
history in the Pacific Northwest. Japanese Journal
of Ecology 32:159–187.
Tumilowicz, J. 1967. Ocena wynikow wprowadzania
niekto- rych obcych gatunkow drzew w lasach
krainy Mazur- sko–Podlaskiej (with Engl.
summary: Evaluation of the results of introduction
of some foreign tree species into the forests of the
Masurian–Podlasie region). Rocznik Den-
drologiczny Polskiego Towarzystwa Botanicznego,
Warszawa (Annals of the Dendrology Section of the
Botanical Society of Poland, Warsaw) 21:136–169.
Turner, J. 1979. Interactions of sulfur with nitrogen in
forest stands pp. 116–125 in Proceedings of Forest
Fertilization Conference, ed. SP Gessel,RM
Kenady, WA Atkinson, Contribution Number 40,
College of Forest Resources, University of
Washington, Seattle.
Tusco, FF. 1963. A Study of Variability in Certain
Douglas-fir Populations in British Columbia.
Doctoral dissertation, University of British
Columbia, Vancouver.
Tyniecki, W. 1891. Wyniki dotychozasowych prob
aklima- tyzacyi obcych drzew w Europie ze
szczegolnym uw- zglednieniem naszego kraju.
Sylwan 35:383–390, 423–425.
Tzschacksch, O. 1981. Stand und Perspektiven der
forstlichen Rauchresistenzzüchtung in der DDR.
Beiträge für die Forstwirtschaft 15(3/4): 134–137.
Tzschacksch, O. 1982. Untersuchungen zur Erblichkeit
der SO2-Resistenz bei Kiefer (Pinus silvestris L.)
und Douglasie (Pseudotsuga menaiesii (Mirb.)
Franco) mit Schlussfolgerun-
gen für die Forstwirtschaft. Beiträge für die Forstwirtschaft
16(3):103–106.
Unestam, T, and E Damm. 1993. Biological control of
seedling diseases by ectomycorrhizae, pp. 173–178
in Diseases and Insects in Forest Nurseries INRA.
Les Colloques, No. 68.
Urbas, J. 1926. Eksote v gozdnem gospodarstvu
slovenije, pp. 363–371 in Pola Stoljea Summarstva
1876–1926 (with French title: Un demisiècle
d’activité forestière 1876–1926). Uredio Aleksandar
Ugrenovi, Zagreb.
USDA Forest Service. 1948. Woody-Plant Seed
Manual. Mis- cellaneous Publication No. 654.
Beltsville Branch, US Department of Agriculture,
US Government Printing Office, Washington, DC.
USDA Forest Service. 1957. California Forest and Range Experi-
ment Station Annual Report 1957, p. 9.
USDA Forest Service. 1958. Timber Resources for
America’s Future. Forest Resource Report No. 14.
342 Douglas-fir: The Genus Pseudotsuga
USDA Forest Service. 1982. An Analysis of the Timber applied at planting. New Forests 2:89–110.
Situation in the United States, 1952–2030. Forest
Resources Report FRR–23.
USDA Western Forest Tree Seed Council. 1973. Tree Seed
Zone Map. Portland (OR): USDA Forest Service in
cooperation with the Western Forest Tree Seed
Council.
US Department of Commerce, Weather Bureau. 1957.
Climatic summary of the United States. Reprinted Series
11. U.S. Department of Commerce, Washington, DC.
US Dept. of Commerce. 1968. Climatic Atlas of the United
States.
Environmental Data Service. 80 p.
US Department of the Interior, Fish and Wildlife Service.
1981. Survey of forest-animal damage in southwest
Oregon. Job Completion Report. Denver Wildlife
Research Center, Forest Animal Research, Denver, CO.
31 p.
Van den Driessche, R. 1969a. Measurement of frost hardiness
in two-year-old Douglas-fir seedlings. Canadian Journal
of Plant Science 49(2):159–172.
Van den Driessche, R. 1969b. Influence of moisture supply,
temperature, and light on frost hardiness changes in
Douglas-fir seedlings. Canadian Journal of Botany
47:1765– 1772.
Van den Driessche, R. 1970. Influence of light intensity and
photoperiod on frost-hardiness development in Douglas-
fir seedlings. Canadian Journal of Botany 48(12):2129–
2134.
Van den Driessche, R. 1975. Flushing Response of Douglas-fir
Bud Stochilling and to Different Air Temperatures after
Chill- ing. British Columbia Forest Service Research
Division, Research Note No. 71, Victoria, BC.
Van den Driessche, R. 1977. Survival of coastal and
interior Douglas-fir seedlings after storage at different
temper- atures, and effectiveness of cold storage in
satisfying chilling requirements. Canadian Journal of
Forest Research 7:125–131.
Van den Driessche, R. 1978. Seasonal changes in root growth
capacity and carbohydrates in red pine and white spruce
nursery seedlings, pp. 6–19 in Proceedings of the IUFR0
Symposium on Root Physiology and Symbiosis, ed. A
Rie- dacker and J. Gagnaire-Michard, Nancy, France.
Van den Driessche, R. 1980. Effects of nitrogen and phos-
phorous fertilization on Douglas-fir nursery growth
and survival after out planting. Canadian Journal of
Forest Research 10:65–70.
Van den Driessche, R. 1984a. Relationship between spacing
and nitrogen fertilization of seedlings in the nursery,
seedling mineral nutrition and out planting performance.
Canadian Journal of Forest Research 14(3):431–436.
Van den Driessche, R. 1984b. Response of Douglas-fir
seed- lings to phosphorus fertilization and the
influence of temperature on this response. Plant and
Soil 72:155–169.
Van den Driessche, R. 1984c. Seedling spacing in the
nursery in relation to growth, yield, and performance
of stock. Forestry Chronicles 60(6):345–355.
Van den Driessche, R. 1984d. Soil fertility in forest nurseries,
pp. 63–74 in Forest Nursery Manual: Production of Bare
Root Seedlings, ed. ML Duryea and TD Landis, Forest
Research Laboratory, OSU, Corvallis, OR/Martinus
Nijhoff/Dr. W. Junk Publishers, Boston, MA. 385 p.
Van den Driessche, R. 1985. Late season fertilization mineral
nutrient reserves and retranslocation in planted Douglas-
fir Pseudotsuga menziesii (Mirb.) Franco. Seedlings.
Forest Science 31(2):485–496.
Van den Driessche, R. 1987. Importance of current photosyn-
thate to new growth in planted conifer seedlings.
Canadian Journal of Forest Research 17:776–782.
Van den Driessche, R. 1988a. Response of Douglas fir (Pseu-
dotsuga menziesii (Mirb.) Franco) to some fertilizers
Van den Driessche, R. 1988b. Response of Douglas-fir Veen, B. 1951. Herkomstonderzoek van de Douglas in
seedlings to phosphorus fertilization and influence Nederland (with Engl. summary: Provenance research of
of temperature on this response. Plant and Soil the Douglas fir in the Netherlands). Proefschrift
80:155–169. Landbouwhogeschool Wageningen (Doctoral thesis.
Van den Driessche, R. 1988c. Nursery growth of conifer Agricultural University Wageningen.)
seed- lings using fertilizers of different solutions
and application time, and their forest growth.
Canadian Journal of Forest Research 18:172–180.
Van den Driessche, R. 1991. Influence of container
nursery regimes upon drought resistance of
seedlings following planting. I: Survival and
Growth. Canadian Journal of Forest Research
21:555–565.
Van den Driessche, R, and KW Cheung. 1979.
Relationship of stem electrical impedance and
water potential of Douglas- fir seedlings to
survival after cold storage. Forest Science
25:507–517.
Van der Kamp, BJ. 1991. Pathogens as agents of
diversity in forested landscapes. Forest Chronicle
67:353–54.
Van der Kamp, BJ. 1993a. Rate of spread of Armillaria
ostoyae in the central interior of British Columbia.
Canadian Journal of Forest Research 23(6):1239–
1241, 10.1139/x93-156
Van der Kamp, B. 1993b. Forest pathology, prepared material.
Module I. 21 p.
Van der Kamp, BJ. 1994. Lodgepole pine stem diseases
and management of stand density in the British
Columbia interior. The Forestry Chronicle 70(6):
773–779.
Van der Kamp, BJ, and J Worrall. 1990. An unusual
case of winter bud damage in British Columbia
interior conifers. Canadian Journal of Forest
Research 20(10):1640–1647.
Van Eerden, E. 1982. Root form of planted trees, pp.
401–405 in Canadian Containerized Tree Seedling
Symposium, September, 14-16,1981. Toronto, ON, ed.
JB Scarratt, C Glerum, and CA Plexman. COJFRC
Symposium Proceedings 0-P-1O, Environment
Canada, Canadian Forestry Service.
Van Eerden, E. 1994. You pay for what you see, p. 67 in
Ab- stracts, Making the Grade, IUFRO Groups s.302–
03, s1.05–04, s2.01–00. ed. DS Maki, TM
McDonough, and TL Noland, St. Marie, Ontario. 45
p.
Van Eerden, E, and J Kinghorn, eds. 1978. Proceedings
of the Root Form of Planting Trees Symposium,
Victoria, BC, May 16-19, 1978. Canadian Forestry
Service Joint Report No. 8, British Columbia
Ministry of Forests, Victoria, BC. 354 p.
Van Haverbeke, DD. 1987. Genetic Variation in
Douglas-fir: A 20–year Test of Provenances in
Eastern Nebraska. USDA Forest Service, General
Technical Report RM–141.
Van Hooser, DD, KL Waddell, JR Mills, and RP
Tymcio. 1991. The interior Douglas-fir resource:
current status and projections to the year 2040, pp.
9–14 in Sympo- sium Proceedings: Interior
Douglas-fir, the species and its management,
February 27–March 1, 1990, comp. and ed. DM
Baumgartner and JE Lotan. Cooperative Extension,
Washington State University, Pullman, WA.
Van Vrendenburch, CLH, and JGA LaBastide. 1969.
The influence of meteorological factors on the
cone crop of Douglas-fir in the Netherlands.
Silvae Genetica 18(5–6) 182–186.
Vapaavuori, EM, R Rikala, and A Ryyppö. 1992.
Effects of root temperature on growth and
photosynthesis in co- nifer seedlings during shoot
elongation. Tree Physiology 10:217–230.
Vargas-Hernandez, J, and WT Adams. 1992. Age-age
cor- relations and early selection for wood, density
in young coastal Douglas-fir. Forest Science
38(2):467–478.
Vazeilles, M. 1946. Les exotiques et la grêle. Bulletin
Société forestière de Franche-Comté 25(12):453.
References 343
Wachtl, FA. 1893. Ein neuer Megastigmus als Samenverwüster
Vega, C. 1990. Ten years’ results of Douglas-fir IUFRO col- von Pseudotsuga douglasii Carr. Wiener Entomologische
lection in North Spain. Paper 2.352 in Proceedings Joint Zeitung 12:24–28, Tab. 1.
Meeting of Western Forest Genetics Association (WFGA)
and IUFRO Working Parties S2.02–05, 06, 12, and 14,
August
20–24, 1990, Olympia, Washington. Weyerhaeuser Co.,
Centralia, WA.
Vegis, A. 1964. Dormancy in higher in plants. Annual Review
of Plant Physiology 15:185 - 224.
Vehov, NK, and VN Vehov. 1962. Hvojnye porody Leso-
stepnoi stancii (Coniferous species at the Soviet
Forest– Steppe Experiment Station). Izolatel’stvo
Ministerstva Kommunal’nogo Hozjajstva RSFSTR,
Moscow. Original not seen, cited from Forestry
Abstracts 25:2045, 1964.
Vertucci, CW and AC Leopold. 1987. Water binding in
legume seeds. Plant Physiology 85:224–231.
Villiers, TA. 1972. Seed dormancy, pp. 219–261 in Seed
Biology Vol. II. ed. TT Kozlowski, Academic Press,
New York. 447 p.
Visart, A de, and C Bommer. 1909. Rapport Sur
l’introduction Des Essences Exotiques en Belgique. Ch.
Buelens, Brussels.
Vogt, KA, RL Edmonds, CC Grier, and SR Piper. 1980.
Seasonal changes in mycorrhizal and fibrous-textured
root mass in 23 and 180-year-old Pacific silver fir stands
in Western Washington. Canadian Journal of Forest
Research 10:523–529.
Volk, K. 1959. Die grüne Douglasie im Forstbezirk
Kandern.
Allgemeine Forstzeitschrift 14(8):181–183.
Volk, K. 1969. Herkünfte, Standortansprüche, waldbauli-
che Behandlung, Leistung und Verwertung der Doug-
lasie aus südbadischer Sicht. Allgemeine
Forstzeitschrift 24(37):715–717.
Volney, WJA. 1984. Competition and survival among insects
colonizing Douglas-fir (Pseudotsuga menziesii [Mirb.]
Franco) cones, pp. 77–84 in Proceedings of the IUFRO
Cone and Seed Insects Working Party Conference.
Working Party S2.07–01. July 31–August 6, 1983,
Athens, GA, ed. HO Yates,
III. USDA Forest Service Southeastern Forest
Experiment
Station, Asheville, NC. 214 p.
Von Aderkas P, and C Leary. 1999. Micropylar exudates
in Douglas fir—timing and volume of production.
Sexual Plant Reproduction 11:354–356.
Von der Gonna, M, and DP Lavender. 1989. Root egress of
white spruce (Picea glauca) seedlings after outplanting
as affected by patch and mound site preparation, pp. 76–
80 in Learning from the Past Looking at the Future, FRDA
Report 030 140, ed. A Mackinnon, and BA Schivenor,
British Columbia Ministry of Forests, Canadian Forest
Service.
von Rudloff, E. 1972. Chemosystematic studies in the genus
Pseudotsuga. I. Leaf oil analysis of the coastal and
Rocky Mountain varieties of the Douglas fir. Canadian
Journal of Botany 50(5):1025–1040.
von Rudloff, E. 1973a. Chemosystematic studies in the
genus Pseudotsuga. II. Geographical variation in the
terpene composition of the leaf oil of Douglas-fir. Pure
and Applied Chemistry 34:401–410.
von Rudloff, E. 1973b. Chemosystematic studies in the genus
Pseudotsuga. III. Population differences in British
Colum- bia as determined by volatile leaf oil analysis.
Canadian Journal of Forest Research 3(3):443–452.
von Rudloff, E. 1975. Volatile leaf oil analysis in chemosys-
tematic studies of North American conifers. Biochemical
Systematics and Ecology 2:131–167.
von Rudloff, E, and GE Rehfeldt. 1980. Chemosystematic
studies in the genus Pseutotsuga. IV. Inheritance and
geographical variation in the leaf oil terpenes of
Douglas- fir from the Pacific Northwest. Canadian
Journal of Botany 58(5):546–556.
Wareing, PF, and NG Smith. 1962. Physiological Studies on
Waddell, KL, DD Oswald, and DS Powell. 1989. Forest the Rooting of Cuttings. Report on Forest Research,
Sta- tistics of the United States, 1987. USDA, Forest 1961/2. Forestry Commission, London.
Service, PNW Research Station, Resource Bulletin
PNW-RB-168.
Wakeley, PC. 1949. Physiological grades of southern
pine nursery stock. Proceedings of the Society of
American Forest- ers 1948:311–322.
Wallin, DO, FJ Swanson, B Manks, JH Cissel, and J
Kertis. 1996. Comparison of managed and
presettlement dynamics in forests of the Pacific
Northwest, USA. Forest Ecology and Management
85:291–309.
Wallis, GW. 1976. Phellinus (Poria) weirii root rot.
Forestry Technical Report 12. Environment
Canada. 15 p.
Wallis, GW, JN Godfrey, and HA Richmond. 1974.
Losses in Fire-killed Timber. Pacific Forest Research
Center, Victoria, BC. 11 p.
Walstad, JD. 1992. History of the development, use, and
management of forest resources, pp. 27–46 in
Reforestation Practices in Southwestern Oregon and
Northern California, ed. SD Hobbs, SD Tesch, PW
Owston, RE Stewart, JC Tappeiner II, and GE
Wells. Forest Research Laboratory, Oregon State
University, Corvallis. 463 p.
Walters, J, and J Soos. 1963. Shoot growth patterns of
some British Columbia conifers. Forest Science
9(1):73–85.
Walther. 1911. Anbau fremdländischer Holzarten.
Allgemeine Forst– und Jagdzeitung 87:154–167.
Wample, RL, RC Durley, and RP Pharis. 1975.
Metabolism of gibberellin A4 by vegetative shoots
of Douglas-fir at three stages of ontogeny.
Physiologia Plantarum 35:273–278.
Wang, CW. 1961. The Forests of China with a Survey of
Grassland and Desert Vegetation. Maria Moore
Cabot Foundation Publication 5. Harvard Universe
Press, Cambridge, MA.
Wareing PF. 1956. Photoperiodism in woody plants.
Annual Review of Plant Physiology 7:191–214.
Wareing, PF. 1958. Interaction between indole-acetic
acid and gibberellic acid in cambial activity.
Nature (London) 181:1744–1745.
Wareing, PF. 1959. Problems of juvenility and
flowering in trees. The Journal of the Linnean
Society of London, Botany 56(366):282–289.
Wareing PF. 1961. Juvenility and induction of flowering.
Recent Advances in Botany 2:1652–1654.
Wareing, PF. 1965. Ecological Aspects of Seed and
Dormancy and Germination, pp. 103–121 in BSBI
conference reports No. 9. Reproductive Biology and
Taxonomy of Vascular Plants. The Botanical Society
of British Isles, Birmingham Uni- versity,
Pergamon Press, New York.
Wareing, PF. 1969. The Control of Bud Dormancy in
Seed Plants. Reprinted from Dormancy & Survival,
XXIIIrd Symposium of The Society for Experimental
Biology at East Anglia University 23:241–262.
Wareing, PF. 1970 Growth and its co-ordination in
trees. In Physiology of Tree Crops. Ed. EC
Luckwill and CW Cutting, Academic Press, London
and NY.
Wareing, PF. 1987. Juvenility and cell determination,
pp. 83–92 in Manipulation of Flowering, ed. JG
Atherton. But- terworth, London.
Waring, RH, and DB Cleary. 1967. Plant moisture stress evalu-
ation by pressure bomb. Science 155:1248, 1253–1254.
Wareing, PF, and VM Frydman. 1976. General aspects
of phase change, with special references to Hedera
helix L. Acta Horticulturae 56:63–69.
Wareing, PF, and IDJ Phillips. 1970. Control of Growth and
Differentiation in Plants. Pergamon Press, Oxford, UK.
Wareing, PF, and PF Saunders. 1971 Hormones and dormancy.
Annual Review of Plant Physiology, 22:261–288.
344 Douglas-fir: The Genus Pseudotsuga
Wareing, PF, and NG Smith. 1963. Physiological Studies on class structure and vertical foliage distribution in Douglas-
the Rooting of Cuttings. Report on Forest Research, fir
1962/63. Forestry Commission, London.
Waring, RH, and ET Youngberg. 1972. Evaluating forest
sites for potential growth response to fertilizer.
Northwest Sci- ence 46(1):67–75.
Webb, WL. 1975. Dynamics of photoassimilated carbon in
Douglas-fir seedlings. Plant Physiology 56:455–459.
Webb, WL. 1977. Seasonal allocation of photoassimilated
car- bon in Douglas-fir seedlings. Plant Physiology
60:320–322.
Webb, WL, and KJ Kilpatrick. 1993. Starch content in Doug-
las-fir diurnal and seasonal dynamics. Forest Science
39(2):359–367.
Webb, WL. 1978. Effects of defoliation and tree energetics,
pp. 77–81 in The Douglas-fir Tussock Moth: A Synthesis,
ed. MH Brookes, RW Stark, and RW Campbell. USDA
Forest Service, Science and Education Agency,
Technical Bulletin 1585. Washington, DC. 331 p.
Webber, BD. 1977. Biomass and nutrient distribution patterns
in a young Pseudotsuga menziesii ecosystem. Canadian
Journal of Forest Research 7:326–334.
Webber, JE. 1987. Increasing seed yield and genetic
efficiency in Douglas-fir seed orchards through pollen
management. Forest Ecology and Management 19:209–
2188.
Webber, JE. 1995. Pollen management for intensive seed or-
chard production. Tree Physiology 15:507–514.
Webber JE, and RA Painter. 1996. Douglas-fir Pollen
Management Manual 2nd ed. Ministry of Forests
Research Program, British Columbia, pp. 9–29.
Webber, JE, and M Stoehr. 1998. Can seed orchards meet
demands for high-gain seed production: Are orchards
obsolete? B.C. Ministry of Forest Research Branch.
In- ternal Report.
Webber, JE, ML Laver, JB Zaerr, and DP Lavender. 1979.
Seasonal variation of abscisic acid in the dormant shoots
of Douglas-fir. Canadian Journal of Botany 57(5):534–
538.
Webber, JE, SD Ross, RP Pharis, and JN Owens. 1985.
Interac- tion between gibberellin A4/7 and root-pruning
on the reproductive and vegetative process in Douglas-
fir. II. Effects on shoot elongation and its relationship to
flow- ering. Canadian Journal of Forest Research 15(2):
348–353.
Weber, W. 1953. Posibilidades económicas de algunas
espe- cies forestales exóticas en Chile. Chile Maderero
3(3):2–3.
Weber, W, and H Gothe. 1954. Naturwaldgesellschaften–
ein forstlicher Briefwechsel. Zeitschrift für
Weltforstwirtschaft 17(6):215–219.
Weck, H. 1949. Fremdländische Holzarten bei der
Wiederherstellung des Waldes Merkblätter des
Zentralinstitutes für Forst– und Holzwirtschaft No. 9.
Weidenbach, P. 1980. Die Douglasie in Baden–
Württemberg– waldbauliche Zielsetzung und zukünftige
Produktion. Holz–Zentralblatt 106(18):277–279.
Weisberg, PJ, and FJ Swanson. 2001. Fire dating from tree
rings in Western Cascades Douglas-fir forests: An error
analysis. Northwest Science 75:145-156.
Weisberg, PJ, and FJ Swanson. 2003. Regional synchroneity
in fire regimes of western Oregon and Washington, USA.
Forest Ecology and Management 172:17–28.
Weise. 1882. Das Vorkommen gewisser fremdländischer
Hol- zarten in Deutschland. Zeitschrift fuer Forst– und
Jagdwesen 14(2):81–103, (3):145–164.
Weiskittel, AR, and DA Maguire. 2007. Response of
Douglas- fir leaf area index and litter fall dynamics to
Swiss needle cast in north Coastal Oregon. Annals of
Forest Science 64:121–132.
Weiskittel, AR, DA Maguire, BA Garber, and A Kanaskie.
2006. Influence of Swiss needle cast on foliage age-
plantations in north coastal Oregon. Canadian Journal Williams, CB, Jr. 1966. Snow damage to coniferous seedlings
of Forest Research 36(6):1497–1508. and saplings. USDA Forest Service, PNW Forest and
Wells, G. 2000. People: Mike Amaranthus. Focus on Forestry Range Experiment Station, Research Note PNW-40. 10
13(1) 22–24. p.
Wells, SP. 1979. Chilling Requirements for Optimal Williams, CG. 2009. Conifer Reproductive Biology. Springer,
Growth of Rocky Mountain Douglas-fir seedlings.
USDA Forest Service Research Note INT-254,
Ogden, UT.
Westwood, MN, and HO Bjornstad.1978. Winter
rainfall re- duces rest period in apple and pear.
Journal of the American Society for Horticultural
Science 103:142–144.
Wetmore, RH. 1943. Leaf-stem relationships in the vascular
plants. Torreya 43:16–28.
Wetzel, SA, and RW Fonda. 2000. Fire history of
Douglas-fir forests in the Morse Creek drainage
of Olympic National Park, Washington.
Northwest Science. 74(4):263–279.
Wheat, J, and RR Silen. 1977. Cooperative tree
improvement in the Douglas-fir region. Third
World Consultation on Forest Tree Breeding
Canberra, Aus. USDA Forest Service.
White, TL. 1987. Drought tolerance of southwestern
Oregon Douglas-fir. Forest Science 33(2):283–293.
White, TL, and KK Ching. 1985. Provenance study of
Douglas- fir in the Pacific Northwest region. IV.
Field performance at age 25 years. Silvae Genetica
34(2/3):84–90.
White, TL, KK Ching, and J Walters. 1979. Effects of
prov- enance, years, and planting location on bud
burst of Douglas-fir. Forest Science 25(1):161–167.
White, TL, DP Lavender, KK Ching, and P Hinz.
1981. First- year height growth of southwestern
Oregon Douglas-fir in three test environments.
Silvae Genetica 30(6):173–178.
Whitehead, DR. 1983, Wind pollination: Some
ecological and evolutionary perspectives, pp. 97–
107 in Pollination Biol- ogy, ed. L Real, Academic
Press, London.
Wheeler, NC, CJ Masters, SC Cade, SD Ross, JW
Keeley, and and LY Hsin. 1985. Girdling: an
effective and practical treatment for enhancing seed
yields in Douglas-fir seed orchards. Canadian Journal
of Forest Research 15(3): 505–510.
Wilamowitz-Möllendorf, von. 1907. Resultate 35-
jähriger Anbauversuche mit ausländischen
Gehölzen, speziell Coniferen, in Gadow.
Mitteilungen der Deutschen Den- drologischen
Gesellschaft 16:135–147.
Wilcox, H. 1962. Growth studies of the root of incense
cedar, Libocedrus decurrens II Morphological
features of the root system and growth behavior.
American Journal of Botany 49:237–245.
Wilcox, MD. 1974. Douglas-fir Provenance Variation
and Selection in New Zealand. New Zealand
Forest Service, Genetics and Tree Improvement
Report No. 69 (unpub- lished), Forest Research
Institute, Rotorua.
Wilcox, MD. 1978. Genetic resources and supply of
douglas fir seed in New Zealand. p 210–217 in A
Review of Douglas- fir in New Zealand. FRI
Symposium No. 15. RN James, and EH Bunn eds.
Forest Service, Forest Research Institute, Rotorua.
Wilen, RW, PF Fu, AJ Robertson, SR Abrams, NH
Low, and LV Gusta. 1996. An abscisic acid analog
inhibits abscisic acid-induced freezing tolerance and
protein accumula- tion, but not abscisic acid-
induced sucrose uptake in a bromegrass (Bromus
inermis Leyss) cell culture. Planta 198:138–143.
Wiley, KN. 1960. Rehabilitation trial of a freeze-
damaged Douglas-fir plantation Weyerhaeuser
Company Forest Research Note 27. 7 p.
Wilhem, S. 1966. Chemical Treatments and the
inoculum potential of soil. Annual Review of
Phytopathology 4:55–78.
References 345
Ministry of Forests and Range Forest Science Program,
New York. Extention Note 101, Victoria, BC. 4 p.
Williamson, RL, and AD Twombly. 1983. Pacific Douglas-
fir, pp. 9–12 in Silvicultural systems for the major forest
types of the United States, RM Burns, comp. US
Department of Agriculture, Agriculture Handbook 445.
Washington, DC.
Willis, CP. 1914. The control of rodents in field seeding. Pro-
ceedings of the Society of American Foresters 9(3):365–
379.
Willis, CP. 1917. Incidental results of a study of Douglas-fir
seed. Journal of Forestry 15(3):991–1002.
Willis, CP, and JV Hofmann. 1915. A study of Douglas-fir
seed.
Proceedings of the Society of American Foresters 10(2):141–
164.
Willson, MF, and N Burley. 1983. Mate Choice in Plants:
Tactics, Mechanism and Consequences. Princeton
University Press, Princeton, NJ.
Wilson, BF. 1970. Evidence for injury as a cause of tree root
branching. Canadian Journal of Botany 48:1497–1498.
Wilson, DO. 1995. The Storage of orthodox seeds, pp. 173–
207 in Seed Quality Basic Mechanisms and Agricultural
Implica- tions, ed. AS Basra, Food Products Press
(Haworth Press, Inc.), New York. 389 p.
Wilson, EH. 1926. Taxads and conifers of Yunnan. Journal
of the Arnold Arboretum 7:37–68.
Wimmer, E. 1909. Anbauversuche mit fremdländischen
Holzarten in den Waldungen des Grossherzogtums Baden.
Paul Parey, Berlin.
Winjum, JK, and NE Johnson. 1962. Estimating Cone Crops
on Young Douglas-fir. Weyerhaeuser Company Forest
Research Note No. 46. 12 p. Centralia, Washington.
Winn, AA. 1988. Ecological and Evolutionary Consequences
of Seed Size in Prunella vulgaris. Ecology 69:1537–
1544.
Wisniewski, M, LH Fuchigatni, JJ Sauter, A Shirazi, and L
Zhen. 1996. Near-lethal stress and bud dormancy in
woody plants, pp. 201–210 in Plant Dormancy.
Physiology, Biochemistry atid Molecular Biology, ed.
GA Lang. CAB International, Wallingford, UK.
Wodehouse, RP. 1935. Pollen grains, their structure,
identification, and significance in science and medicine.
McGraw-Hill Book Company, New York. 574 p.
Reprinted, 1959. Hafner, New York.
Wolfe, JA. 1969. Neogene floristic and vegetational history
of the Pacific Northwest. Madroño 20(3):83–110.
Wood, RF. 1955. The use in Great Brittan of certain
Northwest
American species. Empire Forestry Review 34(3):247–251.
Wood, RF, R Lines, and JR Aldhous. 1960. Provenance stud-
ies: Douglas-fir, pp. 52–53 in Forestry Commission
Report Forestry Research Ended 1959. HMSO, London.
Woodman, JN. n.d. Effect of management practices on
physi- ological processes in forest trees. Unpublished.
Weyer- haeuser Company, Centralia Wa.
Woodman, JN. 1971. Variation of net photosynthesis
within the crown of a large forest grown conifer.
Photosynthetica 5(1):50-54.
Woodruff, DR, BJ Bond, and FC Meinzer. 2004. Does turgor
limit growth in tall trees? Plant, Cell, and Environment
27:229–237.
Woods, JH. Stem girdling to increase seed and pollen
production of coast Douglas-fir. Research Note 103.
British Columbia Ministry of Forests, Victoria, BC.
Woods, JH. 1993. Breeding programs and strategies for
Doug- las-fir in North America. pp. A-X in Breeding
Strategies for Important Tree Species in Canada.
Forestry Canada, Public Report M-X-186E.
Woods, JH, JN King, and CV Cartright. 1995. Early
genetic gains verified in realized-gain trials of coastal
Douglas- fir and western hemlock. British Columbia
NC. 214 p.
Worrall, J. 1971. Absence of “rest” in the cambium of Yates, HO III. 1989. Natural enemies of cone and seed
Douglas- fir. Canadian Journal of Forest Research insects of world conifers, pp. 82–90 in Proceedings of
1:84–89. the 3rd Cone
Worthington, NP. 1961. Some observations on yield and
early thinning in a Douglas-fir plantation. Journal
of Forestry 59(5/May):331–334.
Wraith, JM, and CK Wright. 1998. Soil water and root growth.
HortScience 33(6).
Wright JW. 1952. Pollen Dispersion of Some Forest Trees.
USDA Forest Service, Northeastern Forest
Experiment Station, Paper 46. Radnor, PA.
Wright, AJ, H Knight, and MR Knight. 2002.
Mechanically stimulated TCH3 Gene expression in
Arabidopsis involves protein phosphorylation and
EIN6 downstream of cal- cium. Plant Physiology
128(4):1402–1409.
Wright, E and RF Tarrant. 1958. Occurrence of
Mycorrhizae After Logging and Slash Burning in
the Douglas-fir Type. USDA Forest Service Pacific
Northwest Forest and Range Experiment Station,
Forest Research Note PNW-160, Corvallis, OR. 7
p.
Wright, JW, FH Kung, RA Read, WA Lemmien, and JN
Bright. 1971. Genetic variation in Rocky Mountain
Douglas-fir. Silvae Genetica 20(3):54–60.
Wright, JW. 1953. Pollen dispersion studies: some
practical applications. Journal of Forestry 51:114–
118.
Wulff, E. 1926. Der Nikitsky Botanische Garten in der
Krim. Mitteilungen der Deutschen Dendrologischen
Gesellschaft 37:98–104.
Wyant, JG, PN Omi, and RD Laven. 1986. Fire induced
tree mortality in a Colorado ponderosa
pine/Douglas-fir stand. Forest Science 32:49–59.
Yacubson, D. 1973. The problem of forest seeds in
Argentina. Paper No. 28 in IUFRO Working Party
S2.01–06 Seed Problems, Proceedings of
International Symposium on Seed Processing, Vol. II.
Yamaguchi, DK. 1986. The Development of Old-growth
Douglas-fir Forests Northeast of Mount St. Helens,
Washington, Following an A.D. 1480 Eruption. PhD,
University of Washington, Seattle.
Yanchuk, AD. 1996. General and specific combining
ability from disconnected partial diallels of coastal
Douglas-fir. Silvae Genetica 45(1):37–45.
Yang, JC. 1976. Isoenzyme variation of coastal Douglas-
fir (Pseudotsuga menziesii (Mirb.) Franco var.
menziesii). II. Genotype-environment relationship
in nine provenances. Quarterly Journal of Chinese
Forestry 9:77–89.
Yang, JC. 1978. Provenance trial of Douglas-fir in
Taiwan. Variability of seed weight and seedling
growth. In Chi- nese with English summary.
Quarterly Journal of Chinese Forestry 11(3):1–16.
Yang, JC, TM Ching, and KK Ching. 1977. Isoenzyme
variation of coastal Douglas-fir I. A study of
geographic variation in three enzyme systems.
Silvae Genetica 26(1):10–18.
Yao, C. 1971. Geographic Variation in Seed Weight,
Some Cone Scale Measurements and Seed
Germination of Douglas-fir (Pseudotsuga menziesii
(Mirb.) Franco). MF thesis. Univer- sity of British
Columbia, Faculty of Forestry, Vancouver.
Yapaavuori, EM, and R Rikala. Effects of root
temperature on growth and photosynthesis in
conifer seedlings during shoot elongation. Tree
Physiology 10:217–230.
Yates, HO III. 1984. Cone and seed insects of world
conifers: a comparison with North American fauna,
pp. 26–38 in Proceedings of the IUFRO Cone and
Seed Insects Working Party Conference. Working
Party S2.07–01. July 31–August 6, 1983, Athens, GA,
ed. HO Yates, III. USDA Forest Ser- vice
Southeastern Forest Experiment Station, Asheville,
346 Douglas-fir: The Genus Pseudotsuga
and Seed Insects Working Party Conference, Working man, and M Griffith. 2003. Endochitonase activity in the
Party S2.07-01, IUFRO, 26–30 June 1988, Victoria, BC. apoplastic fluid of Phellinus weirrii-infected Douglas-fir
GE Miller, comp. Forestry Canada, Pacific Forestry and its association with over wintering and antifreeze
Research Centre, Victoria, BC. 242 p. activity. Forest Pathology 33:299–316.
Ye, TZ, KJS Jayawickrama, and JB St. Clair. 2010. Realized Zander, M. 1980. Woody catalogue-conifers in Scotland.
gains from block-plot coastal Douglas-fir trials in the Scot- tish Forestry 34(1):1–12.
northern Oregon Cascades. Silvae Genetica 59(1):29–39. Zavadil, Z, and A Sika. 1978. Introduction of Douglas-fir in
Yeh, FC. 1981. Analysis of gene diversity in some species of the CSSR, pp. 389–396 in Background Papers and
conifers, pp. 48–52 in Isoenzymes of North American Douglas- fir Provenances. Vol.1: IUFRO Joint Meeting
Forest Trees and Forest Insects. MT Conkle, tech. of Working Parties S2.02–05 Douglas-fir Provenances,
coord. USDA, Forest Service General Technical Report S2.02–06 Lodge- pole Pine Provenances, S2.02–12 Sitka
PSW–48. Spruce Provenances, S2.02–14 Abies provenances.
Yeh, FC, and C Heaman. 1983. Heritabilities and genetic Vancouver, Canada 1978. Brit- ish Columbia Ministry of
and phenotypic correlations for height and diameter in Forests, Information Services Branch, Victoria.
coastal Douglas-fir. Canadian Journal of Forest Zavarin, E, and K Snajberk. 1973. Geographic variability of
Research 12(2):181–185. monoterpenes from cortex of Pseudotsuga menziesii.
Yeh, FC, and D O’Malley. 1980. Enzyme variations in Pure and Applied Chemistry 34:411–434.
natural populations of Douglas-fir (Pseudotsuga Zavarin, E, and K Snajberk. 1975. Pseudotsuga menziesii
menziesii (Mirb.) Franco) from British Columbia. I. chemi- cal races of California and Oregon. Biochemical
Genetic variation pat- terns in coastal populations. Silvae Systematics and Ecology 2:121–129.
Genetica 29(3/4):83–92. Zavarin, E, and K Snajberk. 1976. Geographic differentiation
Ying, CC. 1990. Adaptive variation in Douglas-fir, Sitka of cortical monoterpenoids of Pseudotsuga macrocarpa.
spruce and true fir: a summary of provenance research in Biochemical Systematics and Ecology 4:93–96.
coastal British Columbia (abstract), in Proceedings of Zavarin, E, K Snajberk, and WB Critchfield. 1977. Terpenoid
the Joint Meeting of Western Forest Genetics Association chemosystematic studies of Abies grandis. Biochemical
and IUFRO Working Parties, S2.02-05, 06, 12 and 14. Systematics and Ecology 5:81–93.
Olympia, WA, USA.
Zavitkovski, J, and Ferrell, WK. 1968. Effect of drought
Young, E. 1990. Woody plant root physiology, growth, upon rates of photosynthesis, respiration, and
and development: introduction to the colloquium. transpiration of seedlings of two ecotypes of Douglas-fir.
HortScience 25(3):258–259. Botanical Gazette 129(4):346–350.
Žabka, E. 1946. Lesne Prace Badawcze 25:214–23. Zavitkovski, J, and Ferrell, WK. 1970. Effect of drought
Zacharias, K. 1931. Die Douglasie in Sachsen. Mitteilungen der upon rates of photosynthesis, respiration, and
Deutschen Dendrologischen Gesellschaft 43:161–170. transpiration of seedlings of two ecotypes of Douglas-fir.
Zaerr, JB. 1972. Early detection of dead plant tissue. Canadian II. Two-year-old seedlings. Photosynthetica 4:58–67.
Journal of Forest Research 2(2):105–110. Zederbauer, E. 1919. Über Anbauversuche mit fremdlän-
Zaerr, JB, and M Bonnet-Masimbert. 1987. Cytokinin level dischen Holzarten in Österreich. Centralblatt für das
and flowering in Douglas-fir. Forest Ecology and gesamte Forstwesen 45(7/8):153–169.
Manage- ment 19:115–120. Zeevart, JAD, and RA Creelman. Metabolism and physiology
Zaerr, JB, and HR Holbo. 1976. Measuring short-term of abscisic acid. Annual Review of Plant Physiology and
shoot elongation of Douglas-fir seedlings in relation to Plant Molecular Biology 39:439–473.
increas- ing water potential. Forest Science Zeidler. 1956. Erfahrungen im Douglasienanbau-ein Beispiel
22(3):378–382. aus der Sowjetzone. Der Forst und Holzwirt 11(13):271–
Zaerr, JB, and DP Lavender. 1968. Present and potential uses 273.
of growth regulatory substances in Douglas-fir. Zhang, Y, and TD Schowalter. 1997. Cone and seed insect
Proceed- ings, Western Forestry Nursery Council, management in seed orchards: an example from
Biennial Meeting. pp. 68–77. Douglas- fir. Journal of Forestry 95(3):28–32.
Zaerr, JB, and DP Lavender. 1974. The effects of certain cul- Zhong, H, and TD Schowalter. 1989. Conifer bole utilization
tural and environmental treatments upon the growth of by wood-boring beetles in western Oregon. Canadian
roots of Douglas-fir (Pseudotsuga menziesii [Mirb.] Journal of Forest Research 19:943–947.
Franco) seedlings, pp. 27–32 in International Symposium Zieger, E, Pelz, E, and Hornig, W. 1958. Ergebnisse einer
Ecology and Physiology of Root Growth, Academic- Um- frage über Umfang and Art der Frostschäden des
Verlag, Berlin. 438 p. Winters 1955/56 in den staatlichen
Zaerr, JB, and DP Lavender. 1980. Analysis of plant growth Forstwirtschaftsbetrieben der DDR. With Engl.
substances in relation to seedling and plant growth. New Summary: Answers to a questionnaire on the extent and
Zealand, Journal of Forest Research 10:186–195. nature of the frost damage of the 1955/56 winter in the
Zaerr, JB, DP Lavender, M Newton, and RK Hermann. 1981. state forests of the GDR). Archiv für Forst- wesen
Natural versus artificial regeneration in the Douglas-fir 7(4/5):316–337.
region, pp. 177–186 in Wood Power, New Perspectives Zimmerle, H. 1952. Ertragszahlen fuer grüne Douglasie,
on Forest Usage, ed. JJ Talbot and W Swanson. Japa- ner Laerche und Roteiche in Württemberg.
Pergamon Press, Oxford. 239 p. Mitteilungen der Württembergischen Forstlichen
Zak, B. 1964. Role of mycorrhizae in root disease. Annual Versuchsanstalt 9(2):1–44.
Review of Phytopathology 2:377–392. Zimmermann, MH and CL Brown. 1971. Trees, Structure
Zamani, A, RN Sturrock, AKM Ekramoddoullah, J-J Liu, and Function. Springer-Verlag, New York.
and X Yu. 2004. Gene cloning and tissue expression Zon, R. 1913. Effect of source of seed upon the growth of
analysis of a PR-5 Thaumatin-Like protein in Phellinus Douglas fir. Forestry Quarterly 11:499–502.
weirii-infected Douglas-fir. 2004. Phytopathology
94(11):1235–1243.
Zamani, A, RN Sturrock, AKM Ekramoddoullah, SB Wise-
Index
abiotic environment, adverse factors 117, 238, 241, 243, 244, 245, 271. selection 122
See also air pollution, drought, fire, frost, snow, wind. British Columbia 3, 5, 6, 7, 9, 10, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24,
abscisic acid (ABA) 143, 144, 167, 168, 169, 170, 200 25, 26, 36, 37, 43, 44, 45, 52, 53, 60, 62, 68, 75, 76, 78, 81, 82, 84,
Actebia fennica (black army cutworm) 278 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 100, 101, 102,
AFOCEL - Association forêt cellulose 98, 115, 116 103, 104, 105, 106, 107, 108, 110, 111, 112, 114, 115, 116, 120,
Africa 66, 72 121, 122, 123, 155, 160, 176, 185, 188, 201, 208, 209, 211, 212,
air pollutants 213, 214, 215, 216, 219, 224, 227, 231, 232, 239, 247, 248, 252,
fluorides 271
254, 258, 261, 262, 263, 271, 274, 276, 278, 279, 280, 281
sulfur dioxide (SO2) 45, 271 British Columbia Ministry of Forests (BCMP) 123
Alaska 2, 31 British provenance trials 92, 93. See also Great Britain.
anatomy 1, 130, 142, 144, 152, 157, 185, 197, 225, 236, 273 buds 150, 246, 247, 248
Argentina 66, 72, 116 frost injury of 246, 248
arginine 144 bud break, and geographic location 276
Arizona 5, 10, 12, 16, 19, 20, 22, 33, 75, 81, 96, 98, 101, 102, 231, 259, Bulgaria 60, 102, 103, 110
261 provenance trial in 102
Armillaria solidipes [formerly Armillaria ostoyae] 281 caesia, variety of Pseudotsuga menziesii. 4, 5, 44, 64. See also interior
artificial control, of cone and seed insects and diseases 218 Douglas-fir.
systemic and chemical 218
Ashley strain, provenance. See New Zealand. California 5, 6, 7, 9, 10, 15, 16, 17, 18, 19, 23, 24, 25, 26, 32, 35, 38, 51,
Asia 1, 2, 66, 215, 235 56, 65, 68, 70, 74, 75, 78, 79, 80, 84, 90, 92, 94, 95, 96, 97, 98, 99,
associated forest cover 25 100, 101, 106, 107, 110, 111, 112, 113, 114, 115, 116, 121, 123,
Australia 66, 69, 70, 115, 116, 229 173, 177, 185, 189, 201, 208, 211, 212, 213, 221, 227, 258, 261,
provenance trials in 114 263, 264, 268, 276
Austria 33, 50, 61, 102, 103 California black oak (Quercus kelloggii) 25
provenance trials in 102 canyon live oak (Quercus chrysolepis) 25
auxins 143 Cascade Range 5, 18, 23, 24, 25, 26, 41, 75, 79, 80, 89, 100, 101, 102,
Aureobasidium pullulans 223 104, 105, 111, 115, 261, 264, 265, 268, 269, 274
central and eastern North America 32
Baja California 263 chemical races of Douglas-fir 5, 6, 7. See also Pseudotsuga menziesii.
Baltic States 63 Chile 66, 71, 72
Barbara colfaxiana (Douglas-fir cone moth) 207, 208, 209, 212, 213, 215, Chinese Douglas-firs 28, 30
217, 218, 219, 221 Pseudotsuga brevifolia 2, 30
bark beetle 219, 281, 283 Pseudotsuga forrestii 30
Beaumont strain, provenance. See New Zealand. Pseudotsuga gaussenii 2, 29
Belarus 64 Pseudotsuga sinensis 2, 29, 30
Belgium 33, 46, 96, 260 Pseudotsuga wilsoniana 2, 28, 264
provenance trials in 96 chlorophyll fluorescence 204
big-cone Douglas-fir (Pseudotsuga macrocarpa) 1, 7, 26, 27, 264 Choristoneura occidentalis (western spruce budworm) 208, 216, 278, 279
bigleaf maple (Acer macrophyllum) 25 clearcuts 25, 186, 239, 278
biochemical quick tests, and seed germination 180 climate change
biochemistry, seed, and dormancy breaking 167 and chilling requirements 201
biological control, of cone and seed insects and diseases 221 global warming 177, 201, 215, 224, 280, 283
black army cutworm (Actebia fennica) 278
and dormancy and diseases 177, 201, 215, 224, 280,
Bosnia and Herzegovina 60, 103, 104
283 coastal Douglas-fir
provenance trials in 103
Pseudotsuga menziesii var. menziesii 3, 5, 6, 7, 8, 9, 10, 12, 15, 16, 17,
branches
18, 21, 22, 23, 24, 25, 31, 32, 33, 40, 42, 43, 44, 49, 50, 52, 55, 56,
branch thinning 247, 273
60, 61, 62, 63, 64, 70, 74, 77, 78, 79, 80, 81, 84, 85, 90, 91, 95, 96,
ontogeny of
97, 98, 100, 101, 102, 104, 117, 118, 120, 123, 166, 198, 209, 245,
and disease 280, 281
246, 247, 250, 251, 252, 253, 254, 257, 258, 259, 260, 261, 262,
and growth 273
263, 264, 268, 270, 271, 272, 274
breeding and improvement 65, 80, 82, 89, 91, 117, 122
Pseudotsuga menziesii var. viridis 4
goals and objectives 121
cold acclimation and deacclimation 249
program, steps in 122
cold hardiness
tree breeding/genetic improvement programs
and age 251
North America 122
assessment of 98, 106, 118, 119, 120, 140, 197, 198, 200, 201, 202, 217,
Northwest Tree Improvement Program (NWTIC) 122
227, 250, 251, 252, 253, 254, 255, 256, 257, 258, 259, 261, 262
Weyerhaeuser Company 123
British Columbia Ministry of Forests (BCMP) 123
Inland Empire Tree Improvement Program (IETIC) 123 Note: this brief index is intended only as a supplement for the
Europe 123 limited- edition print version and is not a complete index. Please see
also the detailed table of contents, as well as the electronic version
New Zealand 124
of the entire book, which is fully searchable and freely available
breeding zones 121, 122, 258
online, in ScholarsArchive@OSU, at this address:
http://hdl.handle.net/1957/47168.
347
348 Douglas-fir: The Genus Pseudotsuga
quantitative genetics of cytological studies 1, 2, 5, 197
coastal variety 252 fossil record 1, 2, 3, 10, 15, 31, 236, 306
interior variety 254 Cretaceous 1, 2, 10, 236
fall 98, 106, 118, 119, 120, 140, 197, 198, 200, 201, 202, 217, 227, 250, Pleistocene 3, 4, 7, 9, 15, 19, 28, 31
251, 252, 253, 254, 255, 256, 257, 258, 259, 261, 262 geographic variation 10, 11, 117, 119, 122, 251, 253, 257, 258
spring 98, 106, 118, 119, 120, 140, 197, 198, 200, 201, 202, 217, 227, intraspecific variation 4, 6, 7, 12
250, 251, 252, 253, 254, 255, 256, 257, 258, 259, 261, 262 morphological characteristics 3
cold storage, of seedlings 202, 203 Pseudotsuga, the genus 1
cold water spray, and cone and seed insects 219 differentiation into varieties 3, 6, 12, 13, 14, 79, 254, 255, 256
Colorado 3, 5, 9, 10, 16, 19, 22, 32, 33, 70, 75, 81, 85, 86, 96, 98, 101, Sierra Nevada race 7, 10
105, 107, 108, 110, 137, 211, 259, 261
fertilization. See pollen.
common garden studies
fertilizers (mineral nutrition) 159, 160
variety glauca 81
and flower induction 132
variety menziesii 81
Finland 53, 110, 217
competition, seedling 220, 277
provenance trials in 110
cone and seed
fire
insects 207
and associated forest cover 25
See Douglas-fir cone moth
and deterioration of fire-killed Douglas-fir 270
natural control of 220, 221
effects of, as an abiotic factor in Douglas-fir forests 243, 268
conifer seedling weevil (Steremnius carinatus) 278
frequency, and range of Douglas-fir 18, 25
Contarinia oregonensis (Douglas-fir cone gall midge) 207, 208, 213,
history, prevention, and management of, 25, 26
214, 218, 219, 221
flowering 125, 130
Contarinia washingtonensis (Douglas-fir cone scale midge) 208,
embryogeny. See pollen.
214
history and nomenclature of 125
Cooley spruce gall adelgid (Adleges [Gilletteella] cooleyi) 10, 37,
initiation of 125, 131
39 production 185, 207
induction of
Croatia 56
cultural treatments 137
cultivation, seedling 36, 38
girdling 137
cultural treatments 137
gravimorphism, shading 140
cutting test, seed 180
root pruning and grafting 139
Cyprus 56
top pruning and branch thinning 139
cytokinins 143, 324
methodology of 131, 138, 140, 200
Czech Republic 57, 58, 104
fertilizers (mineral nutrition) 132
provenance trials in 104
light intensity 131
Dendroctonus pseudotsugae (Douglas-fir beetle) 279, 280-281 moisture stress 131
Denmark 33, 50, 51, 53, 83, 86, 107, 110, 151, 245 photoperiod 132
provenance trials in 106, 107 shoots 133
diapause, in cone and seed insects 216, 217 temperature 133, 134
Dioryctria abietella 207, 208, 215, 220 juvenility and maturity and 127
diseases 222, 240, 281. See also fungi; foliage, root, cone and seed plant growth regulators and 140
diseases. abscisic acid 143
control of 135, 218 arginine 144
Mexico 1, 5, 7, 10, 12, 16, 17, 19, 20, 21, 22, 26, 33, 75, 81, 87, 89, 96, 98, auxins 143
99, 101, 102, 105, 110, 114, 211, 259, 261 carbohydrates 144
dormancy. See seed, seedling. cytokinins 143
Douglas, David 17, 32, 33, 35, 36, 40, 41, 45, 51, 54, 94 gibberellins 140
Douglas-fir beetle (Dendroctonus pseudotsugae) 279, 280-281 polyamines 144
Douglas-fir cone moth (Barbara colfaxiana) 207, 208, 209, 212, 213, 215, sex expression 143
217, 218, 219, 221 promotion of 145, 291, 329, 333
Douglas-fir tussock moth (Orgyia pseudotsugata) 278, 279 fluorides 271
drought. See also frost-induced drought. foliage
and cold hardiness 254 anatomy, quantity, distribution 273
and regeneration 265 diseases
-caused damage 62, 266 Rhabdocline needle cast (Rhabdocline pseudotsugae) 31
-caused mortality 60, 61, 72 Swiss needle cast (Phaeocryptopus gaeumannii) 17, 33, 39, 44, 49,
resistance 17, 31, 97, 261, 266 51, 52, 55, 68, 70, 85, 87, 113, 115, 121, 122, 224, 281, 282, 283
Eastern Europe 57 Fort Bragg strain, provenance. See New Zealand.
ecology 239 fossil record 1, 2, 3
ectomycorrhizae 235, 236. See also mycorrhizae. France 33, 47, 48, 49, 66, 96, 97, 98, 99, 115, 116, 144, 211, 265, 276
electrophoresis 7 provenance trials in 96
embryogeny 145. See pollen. AFOCEL - Association forêt cellulose 98, 115,
enzymes 116 recommendations based on 99
allozyme 6, 7, 8, 9, 10, 11, 12, 13, 14, 75, 78, 80, 81, 106, frost
isozymes 7, 116 -induced drought 260, 261
Estonia 64, 110, 111 and site aspect 260
provenance trials in 110 hardiness 33, 81, 90, 118, 120, 245, 249, 252, 254, 257, 258, 259, 262
Europe 4, 31, 32, 33, 34, 38, 40, 48, 50, 54, 57, 58, 65, 83, 89, 100, 108, prediction models 257
110, 112, 115, 116, 122, 123, 177, 186, 207, 211, 212, 215, 220, heaving 52, 264
221, 230, 235, 251, 260, 282 injury
evolutionary history consequences of 33, 45, 245, 246, 248, 249, 250, 251, 252, 260
ancestral Douglas-fir (Eocene Pseudotsuga sonomenis) 3
Index 349

identification of 33, 45, 245, 246 See also interior Douglas-fir.


Frothingham, Earl Hazeltine, and Douglas-fir growing regions 4 global warming. See climate change.
fungi 45, 85, 143, 194, 221, 222, 223, 224, 226, 235, 236, 237, 238, grand fir (Abies grandis) 26, 269
239, gravimorphism 140, 318
240, 241, 242, 243, 244, 248, 283 Great Basin 3, 7
and cone and seed diseases 222, 223 Great Britain 33, 36, 37, 38, 39, 40, 62, 186, 211
Aspergillus spp. 223 Greece 56
Aureobasidium [Pullularia] spp. 223, 224 growth regulators 169
Caloscypha fulgens 223, 224
hail 264
Cephalosporium 224
Hawaii 31
Corticium pini-canadensis 223
height growth 116, 274
Fusarium spp. 140, 223, 224, 240, 281
heritabilities
Gliocladium spp. 223, 224
and amounts of genetic variation 119
Mucor spp. 223
history and nomenclature 125
Penicillium spp. 223 Hungary 59, 60, 86, 110
Spicaria spp. 223 provenance trials in 105
Trichoderma spp. 223, 224 hybridization, intraspecific 167, 262, 263, 264
Trichothecium spp. 224 hydrogen peroxide, and seed germination testing 180
and foliage diseases
Phaeocryptopus gaeumannii (Swiss needle cast) 17, 33, 39, 44, 49, ice 264, 269
51, 52, 55, 68, 70, 85, 87, 113, 115, 121, 122, 224, 281, 282, 283. Idaho 3, 7, 10, 12, 16, 19, 20, 22, 25, 26, 33, 75, 82, 83, 105, 106, 107,
Phellinus weirii (laminated root rot) 139, 265, 281, 282 110, 123, 195, 211, 216, 255, 256, 257, 261, 263, 268, 269, 272,
and root diseases inbreeding 254
Armillaria solidipes 281 incense cedar (Calocedrus decurrens) 25
Phellinus weirii 139, 265, 281, India 66, 215, 302, 320, 343
282 and root and butt rots Inland Empire 122, 259
Polyporus schweinitzii (Schweinitzii root and butt rot) 32, 282, 283 Inland Empire Tree Improvement Program (IETIC) 123
Fomes pini (conk rot) 282, 283 insects
Fomitopsis officinalis (brown trunk rot or quinine fungus) 282, 283 control of
Fomitopsis cajanderi [formerly Fomes subroseus] (yellow brown artificial 218
top systemic and chemical attractants 218
rot) 282 barriers 220, 240
natural, of cone and seed insects 220, 221
genecology of 118 biological 221
estimates of genetic gain 119, 120, 123 competition 220
heritabilities and amounts of genetic variation 119
phenology
genetic correlations 119
barriers 219
quantitative genetics and inheritance 118
cold water spray 220
genetic improvement and tree breeding programs 122. See also
seed treatment 220
breeding. diapause in 216, 217
North America 122 pests of Douglas-fir
Northwest Tree Improvement Program (NWTIC) 122 bark beetle 219, 281, 283
Weyerhaeuser Company 123
black army cutworm (Actebia fennica) 278
British Columbia Ministry of Forests (BCMP) 123
conifer seedling weevil (Steremnius carinatus) 278
Inland Empire Tree Improvement Program (IETIC) 123
Cooley spruce gall adelgid (Adleges [Gilletteella] cooleyi) 10, 37, 39
Europe 123 Douglas-fir beetle (Dendroctonus pseudotsugae) 279, 280-281
New Zealand 124 Douglas-fir cone gall midge (Contarinia oregonensis) 207, 208, 213,
Germany 31, 33, 40, 41, 43, 44, 45, 47, 49, 57, 66, 84, 85, 86, 87, 88, 89, 214, 218, 219, 221
90, 91, 107, 110, 116, 123, 124, 246, 247, 248, 251, 258, 260, 262, Douglas-fir cone moth (Barbara colfaxiana) 207, 208, 209, 212, 213,
264, 271 215, 217, 218, 219, 221
provenance trials in 14, 83 Douglas-fir cone scale midge (Contarinia washingtonensis) 208,
1930s 85 214
Baden-Wuerttemberg (1955/1958) 88 Douglas-fir seed chalcid (Megastigmus spermotrophus) 207, 208,
Fabricius (1930) 87 209, 210, 211, 212, 216, 217, 218, 219, 220, 221
Fürstenberg (1904-1911) 83 Douglas-fir tussock moth (Orgyia pseudotsugata) 278, 279
German Democratic Republic (1961) 89 spruce coneworm (Dioryctria abietella) 215
Hessian Forest Experiment Station (1958) 89 weevil, Lepesoma lecontei 216
Kaiserslautern (1912) 84 western blackheaded budworm (Acleris gloverana) 279
Lower Saxony Forest Experiment Station (1954/1958) 87 western conifer seed bug (Leptoglossus occidentalis) 207, 209-210,
Schmalenbeck Institute of Forest Genetics and Forest Tree 216, 219
Breeding (1962-1963) 89 western hemlock looper (Lambdina fiscellaria lugubrosa) 279, 280
Schober trial (1958) 88 western spruce budworm (Choristoneura occidentalis) 208, 216,
Schwappach (1910) 84 278, 279
Wiedemann (1932–1933) 85 interior (inland) Douglas-fir
germination tests 107, 158, 170, 177, 178, 181, 189, 195, 223. See also Pseudotsuga menziesii var. caesia 4, 5, 44, 64
seed, germination. Pseudotsuga menziesii var. glauca (Rocky Mountain Douglas-fir) 3, 4,
giant chinkapin (Chrysolepis chrysophylla) 25 5, 6, 10, 12, 16, 17, 21, 25, 32, 33, 36, 42, 43, 44, 45, 46, 50, 52,
gibberellins 140, 169, 300, 328, 329, 331 54, 61, 62, 63, 64, 70, 72, 75, 81, 82, 84, 85, 86, 87, 89, 90, 91, 94,
girdling 137, 138, 139, 142, 249 96, 98, 101, 102, 104, 105, 106, 108, 110, 111, 115, 166, 189, 209,
glaciers, glaciation 3, 4, 8 212, 231, 245, 254, 257, 258, 262, 263, 264, 269, 274, 277, 278,
glauca, variety of Pseudotsuga menziesii 3, 4, 5, 6, 10, 16, 17, 32, 33, 44, 280
45, 54, 64, 75, 81, 82, 84, 85, 86, 87, 91, 94, 98, 101, 102, 105,
106, 108, 110, 111, 115, 189, 209, 212, 231, 262, 277, 278, 280.
350 Douglas-fir: The Genus Pseudotsuga
intervarietal hybrids 7, 14, 331 ectomycorrhizae 235
intraspecific variation 4-6, 301 anatomy of 236
and testing inoculation with 237, 238, 239, 242, 244
terpenes 6 physiology 242
enzymes 7 mycorrhizosphere 241
random amplified polymorphic DNA (RAPD) 12 and nutrient and water uptake 241, 261
introduced populations, progeny of 73, 81, 88, 89, 94, 102, 110, 116
natural range 9, 14, 17, 25, 31, 32, 33, 53, 66, 82, 83, 89, 98, 100, 106,
introduction, areas of. See also individual countries and states.
110, 115, 245, 253, 259, 262, 272, 276
Northern Hemisphere 31
altitudinal distribution 21
Asia 66
area occupied by 22
Europe 33
associated forest cover 25
eastern 57-62
climate 23
Baltic States 63-65
history 15
Great Britain and Ireland 33, 39
soils 24
Mediterranean 54-56
natural reproduction, and disease resistence 281
northern 50 needles 16, 245, 246. See also foliage.
western central 40, 45-47, 49, 50 ontogeny, anatomy, quantity, distribution 273
North America Netherlands, The 45, 187
central and eastern 32 provenance trials in 95
western 31 New Mexico 5, 10, 12, 16, 17, 19, 33, 75, 81, 89, 96, 98, 101, 102, 105,
Southern Hemisphere 66 110, 211, 259, 261
Africa 72 New Zealand 66, 67, 68, 69, 112, 113, 116, 124, 137, 211, 245
South America 71, 72 provenance trials in 112
southwestern Pacific 66, 69 Ashley strain 114
Ireland 33, 39, 40, 94, 95, 203 Beaumont strain 114
provenance trials in 94 Fort Bragg strain 114
isozyme study 14 Kaingaroa strain 114
Italy 55, 56, 100, 10 North America 1, 2, 3, 15, 17, 31, 32, 33, 44, 81, 83, 97, 106, 110, 112,
provenance trials in 100 117, 121, 122, 124, 131, 201, 202, 203, 207, 209, 212, 215, 220,
Kaingaroa strain, provenance. See New Zealand. 221, 230, 235, 260, 262, 268, 278, 280, 281
Northern Europe 50
Lambdina fiscellaria lugubrosa (western hemlock looper) 279, 280 Northern Hemisphere 31
laminated root rot (Phellinus weirii) 139, 265, 281, 282 Northwest Tree Improvement Cooperative 120, 122, 252
Larix 1, 2, 26, 39, 46, 147, 148, 152, 155, 161, 162, 230, 269300, 305, 317, Northwest Tree Improvement Program (NWTIC) 122
321, 322, 336 Norway 33, 41, 45, 46, 47, 49, 51, 52, 53, 56, 59, 62, 63, 64, 65, 108, 110,
Latvia 63, 64, 110 115, 145
provenance trials in provenance trials in 108
110 old growth 16, 25, 274, 275, 276, 277, 278, 281, 282, 283. See also
Lepesoma lecontei 207, 216 mature trees.
Leptoglossus occidentalis (western conifer seed bug) 207, 209-210, 216, ontogeny of
219 and disease 281, 282-284
light 131, 172, 187, 203, 209, 233 and growth 273
intensity level 131, 172, 187, 203, 209, 233 and height 274
Lithuania 63, 64, 110 ontogeny, of Douglas-fir
Luxembourg 46, 47 and disease 281-284
mature trees and stands 273, 274, 277, 278, 282, 283. See also old and growth 273-274
growth. and height 274-275
ontogeny of and insects 278-281
and disease 282-284 and phenology 275-276
and growth 273 and photosynthesis 276-278
and height 274 Oregon 3, 5, 6, 7, 9, 10, 12, 14, 16, 18, 19, 21, 22, 23, 24, 25, 26, 32, 35,
and insects 278, 280 37, 38, 41, 43, 49, 50, 51, 56, 61, 68, 70, 72, 73, 74, 75, 78, 79, 80,
and phenology 275-276 81, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100,
and photosynthesis 277 101, 102, 103, 104, 105, 106, 107, 108, 111, 112, 113, 114, 115,
Mediterranean Europe 54 116, 117, 118, 120, 121, 122, 123, 139, 143, 172, 174, 177, 182,
Megastigmus spermotrophus (Douglas-fir seed chalcid) 207, 208, 209, 186, 187, 189, 201, 207, 208, 209, 210, 211, 212, 213, 216, 227,
210, 211, 212, 216, 217, 218, 219, 220, 221 238, 239, 241, 245, 247, 248, 249, 250, 252, 253, 254, 255, 256,
Menzies, Archibald 15, 17 257, 258, 259, 260, 261, 263, 264, 265, 266, 267, 268, 269, 270,
menziesii, variety of Pseudotsuga menziesii 3, 5, 6, 7, 8, 9, 10, 12, 15, 16, 272, 274, 276, 281, 282, 283
17, 18, 21, 22, 23, 24, 25, 31, 32, 33, 40, 42, 43, 44, 49, 50, 52, 55, Oregon white oak (Quercus garryana) 25
56, 60, 61, 62, 63, 64, 70, 74, 77, 78, 79, 80, 81, 84, 85, 90, 91, 95, Organic residues 233
96, 97, 98, 100, 101, 102, 104, 117, 118, 120, 123, 166, 198, 209, Orgyia pseudotsugata (Douglas-fir tussock moth) 278, 279
245, 246, 247, 250, 251, 252, 253, 254, 257, 258, 259, 260, 261, ovulate strobilus 155. See also pollen, embryogeny.
262, 263, 264, 268, 270, 271, 272, 274. See also coastal Pacific madrone (Arbutus menziesii) 25
Douglas-fir. Pacific Northwest provenance trials 73, 74, 75, 78, 81
microsporangiate strobilus 148, 150. See also pollen, embryogeny. Penicillium spp. 223
moisture stress 131, 173, 183, 267 Phaeocryptopus gaeumannii (Swiss needle cast) 17, 33, 39, 44, 49, 51,
mycorrhizae 235-244 52, 55, 68, 70, 85, 87, 113, 115, 121, 122, 224, 281, 282, 283.
and diseases of Douglas-fir 240 Phellinus weirii (laminated root rot) 139, 265, 281, 282
cost-benefits of 243 phenology
ecology 239
Index 351

bud break, and cold water spray 219 Pseudotsuga menziesii, varieties of 15
photoperiod 125 var. caesia. See interior Douglas-fir.
and cold hardiness 249, 258 var. glauca. See interior Douglas-fir.
and dormancy 198, 199, 200, 202 var. menziesii. See coastal Douglas-fir.
and flowering 132 var. viridis. See coastal Douglas-fir.
and germination 173 Pseudotsuga punctata 1
and seedling tests 79 Pseudotsuga sinensis 2, 29, 30
responses to differences in, by variety 5 Pseudotsuga sonomenis 3
Pinaceae 1, 140, 142, 148, 149, 150, 155, 160, 161, 235, 236, 238, 244, Pseudotsuga wilsoniana 2, 28, 264
Pinus spp. 1, 15, 25, 26, 27, 31, 35, 53, 147, 148, 149, 152, 155, 160, 161, Pullularia pullulans [Aureobasidium pullulans] 223
162, 166, 215, 228, 230, 232, 241, 269, 272
quantitative genetics and inheritance 118, 252
plant growth regulators
and dormancy 140, 199 random amplified polymorphic DNA (RAPD) 12, 13, 14, 16
and flowering and phenotype analysis 13
abscisic acid 143 Rhabdocline needle cast (Rhabdocline pseudotsugae) 31
arginine 144 reproduction, natural, and disease resistence 281
auxins 143 reproductive buds 248
carbohydrates 144 Republic of South Africa (South Africa) 72
cytokinins 143 Rocky Mountain Douglas-fir (Pseudotsuga menziesii var. glauca). See
gibberellins 140 interior Douglas-fir.
polyamines 144 Rocky Mountains 3, 7, 16, 17, 21, 22, 23, 24, 25, 26, 32, 33, 82, 216, 254,
sex expression 143 255, 259, 263, 269
Poland 33, 62, 63, 85, 105, 106, 110, 258 Romania 61, 62, 311
provenance trials in 105 root
pollen diseases
distribution 162, 163 Armillaria root (Armillaria solidipes) 281
embryogeny 145 laminated root rot (Phellinus weirii) 139, 265, 281, 282
microsporangiate strobilus 148, 105 pruning and grafting 139
ovulate strobilus 155 systems
pollen cone, the 149, 150 and seedling survival and growth 227
fertilization 159-160 temperature
physiology 141, 148, 149, 150, 157, 158, 160, 161, 162 growth 135, 225, 226, 246, 249
pollination 150, 159, 161, 163 frost injury of 249
storage 157 root and butt rots
viability of 157, 158, 159 Polyporus schweinitzii (Schweinitzii root and butt rot) 32, 282, 283
polyamines 144, 297 Fomes pini (conk rot) 282, 283
ponderosa pine (Pinus ponderosa) 25, 269 Fomitopsis officinalis (brown trunk rot or quinine fungus) 282, 283
Port Orford cedar (Chamaecyparis lawsoniana) 25 Fomitopsis cajanderi [formerly Fomes subroseus] (yellow brown top
Portugal 33, 54 rot) 282
provenance trials 56, 63, 73, 74, 75, 82, 83, 84, 87, 88, 92, 96, 100, 102, root growth periodicity 234
103, 105, 108, 110, 111, 258, 260. See also individual countries. root electrolyte leakage (REL) 229
knowledge gained from 115 root regeneration potential 227, 229
height growth 116 Russia 1, 41, 65, 110
progeny of introduced populations 116
saplings, young trees 270, 273, 274, 276, 281. See also young growth.
survival 115
ontogeny of
Northern Hemisphere
and competition 277-278
Asia 111
and disease 281
Europe
and growth 273
eastern 102, 103, 104, 105
and height 274
Baltic States 110
and insects 278, 280
Great Britain and Ireland 92, 94
and phenology 276
Mediterranean 99, 100, 111
and photosynthesis 276
northern 106, 108, 110
seed
western central 83-89, 95, 96, 98, 102
biochemistry of 167
IUFRO 83, 90
growth regulators and 169
GDR 91
gibberellins 169
Hesse (1970) 90
abscisic acid (ABA) 170
seed sources for Germany 91
collections, IUFRO 5, 64, 78, 80, 88, 95, 99, 111, 122, 124, 185, 258,
Pacific Northwest 73
260
Oregon State University (1954) 74
development 167, 168, 190, 210, 213, 224
British Columbia Forest Service 75
dispersal 14
University of British Columbia 78
dormancy 165, 166, 167, 168, 169, 170, 183, 192, 289, 296, 312, 326,
Short-term seedling tests 78
common garden: variety glauca 81 329, 338
Southern Hemisphere 116 dormancy breaking 167
Southwestern Pacific 112, 114 flight 188
Pseudotsuga brevifolia 2, 30 germination 170
Pseudotsuga forrestii 30 and light 172
Pseudotsuga gaussenii 2, 29 and moisture 173
Pseudotsuga japonica 2, 27, 28 and temperature 175
Pseudotsuga macrocarpa 2, 7, 26, 27 tests 107, 158, 170, 177, 178, 181, 189, 195, 223
352 Douglas-fir: The Genus Pseudotsuga
standard germination test 179 Southern Hemisphere 116
cutting test 180 Southwestern Pacific 66
biochemical quick tests Soviet Union (former) 64, 65
180 Spain 33, 54, 55, 99, 100, 115, 309, 344
hydrogen peroxide 180 provenance trials in 99
excised embryo 181 Sri Lanka 66
seed-vigor tests 181 Stem tissue 248
stratification 166, 174, 175, 182 Steremnius carinatus (conifer seedling weevil) 278
production of, and cones 185 sugar pine (Pinus lambertiana) 25
processing (damage) 193, 288 sulfur dioxide (SO2) 45, 271
size Sweden 33, 52, 53, 108, 109, 110, 262
and anatomy 185 provenance trials in 108
and germination 185, 189 Swiss needle cast (Phaeocryptopus gaeumannii) 17, 33, 39, 44, 49, 51,
storage 168, 176, 181, 193, 194, 195 52, 55, 68, 70, 85, 87, 113, 115, 121, 122, 224, 281, 282, 283
under controlled conditions 194 Switzerland 33, 49, 50, 116, 267
under natural conditions 195
Taiwan, distribution 1, 28, 29, 111
treatment, for insect infestation 220
viability of 176, 180, 193, 194, 195, 220, provenance trial 111
224 tanoak (Lithocarpus densiflorus)
seedlings 5, 32, 60, 78, 79, 81, 84, 89, 90, 91, 94, 102, 109, 112, 118, 120, 25 taxonomic divisions 4
134, 197, 228, 247, 248, 251, 253, 254, 255, 259, 260, 266, 273, temperature, effects of
275, 276, 280, 281 heat. See climate change.
and climate change and chilling 201 cold 133, 175, 177, 199, 230. See also climate change.
cold storage of terpenes 6, 7, 9, 10, 13, 279
and stress resistance 202 tree breeding. See breeding.
with light 203 trunk 246, 339
without light 203 Turkey 56, 111, 337
dormancy 197, 204, 227, 229, 230 provenance trials in 111
and concentrations of inorganic and organic constituents 204 Ukraine 64, 65, 110
and growth potential 202 Utah 5, 10, 12, 16, 19, 26, 96, 98, 101
and chlorophyll fluorescence 204
and physiological response to 199 varieties of Douglas-fir. 3, 15. See also coastal Douglas-fir, interior
and plant growth regulators 199 Douglas-fir.
initiation of, and environment 199 differentiation into 3
dormancy breaking 200 chemical races 5, 6, 7
ontogeny of viridis, variety of Pseudotsuga menziesii 4. See coastal Douglas-fir.
and disease 280, 281 Washington 3, 5, 6, 7, 10, 16, 18, 21, 22, 23, 24, 25, 26, 35, 36, 37, 40, 41,
and growth 273 43, 46, 47, 49, 50, 51, 53, 56, 60, 61, 62, 67, 68, 72, 73, 75, 78, 79,
and height 274 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99,
and insects 278, 280 100, 101, 102, 103, 104, 105, 106, 107, 108, 110, 111, 112, 114,
and phenology 275 115, 116, 117, 118, 121, 123, 136, 187, 201, 208, 209, 211, 213,
and photosynthesis 276 214, 216, 227, 239, 245, 246, 248, 249, 251, 252, 255, 256, 257,
selection of 65, 118, 122, 123 258, 259, 260, 261, 263, 265, 266, 267, 268, 269, 270, 271, 272,
survival 82, 92, 100, 108, 109, 110, 115, 203, 227, 262, 263, 274, 283
tests 75, 78, 79, 122, 228. western central Europe 40
short-term 78. See also common garden and provenance trials. western conifer seed bug (Leptoglossus occidentalis) 207, 209, 210, 216,
Sequoia sempervirens 16 219
Serbia 60 western hemlock looper (Lambdina fiscellaria lugubrosa)
sex expression 143 western larch (Larix occidentalis) 26, 269
shading Western North America 31
shade 132, 140, 233, 277 western spruce budworm (Choristoneura occidentalis) 208, 216, 278, 279
shoots 133 Weyerhaeuser Company 122, 123, 131
effects of temperature on 133, 134 What has been learned from provenance experiments? 115
growth and development of 127, 128, 129, 130, 246 white fir (Abies concolor) 25
female reproductive 125 wildfire 25, 268, 270. See also fire.
formation of 131 wind
injury to 246, 247, 248, 249, 250, 252, 253, 254, 256, 258 and seed flight 188, 189
lammas 131, 132, 134 damage from (windthrow) 25, 272, 283
Sierra Nevada 3, 6, 7, 10, 19, 21, 23, 25, 70, 97, 112 effect of, on cone crop 186
Sierra Nevada population 10 pollination 79, 80, 106, 118, 119, 122, 163, 164, 263
Slovakia 58, 59, 105 Wind River area, nursery and study sites 74, 239, 246, 251, 273, 278
provenance trials in 105
Slovenia 56 young growth, trees and stands 260, 273, 277, 281, 282. See also
snow 264, 265 saplings.
soil temperatures 134, 135, 137, 176, 185, 230, 231, 232, 233, 239, 242
South America 71
Douglas-fir, the genus Pseudotsuga, has a wide distribution in North America, and
includes some of the most widely distributed species outside of their natural ranges.
Douglas-fir has been introduced to Europe, New Zealand, South America, and
elsewhere around the world. It is an accepted and integral part of forest management
in many countries because of its economic importance and its reputation as a tree that
may be able to deal with climate change.
This book provides an overview of research activities and findings that highlight unique
aspects of Douglas-fir physiology, genetics, and other related issues. The authors have
pulled together a tremendous amount of information, beginning with the evolutionary
history and distribution of Douglas-fir. They provide detailed descriptions of the
introductions of Douglas-fir to other countries, including information about initial
plantings, provenance trials, and genetic tree improvement activities. In sections about
life history, the authors bring to bear their long-time research and teaching
experiences, as well as detailed descriptions of flowering, seed, root, and seedling
physiology, followed by sections on biotic factors, such as mycorrhizae, insects, and
diseases, and abiotic factors, including frost, drought, and fire.
Douglas-fir: The Genus Pseudotsuga will stand the test of time as an invaluable
collection and cornerstone of information that could have easily been lost as
researchers, educators, and managers are inundated with new research results. It is
intended as a resource for everyone interested in understanding the opportunities and
challenges of managing Douglas-fir in a variety of regions and settings. It provides
information for historians and social scientists investigating forestry trends;
researchers, educators, and managers looking for detailed information in areas such as
genetics and regeneration practices; and all others interested in the beautiful trees we

ISBN 978-0-615-97995-3
Oregon Forest Research Laboratory
College of Forestry
Oregon State University
Corvallis, Oregon

You might also like