24.02.2013 Views

View or print this publication - Northern Research Station - USDA ...

View or print this publication - Northern Research Station - USDA ...

View or print this publication - Northern Research Station - USDA ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

"


ABOUTTHE COVER<br />

On the cover are artistic renditions of five f<strong>or</strong>est insects and/<strong>or</strong> their damage. None are drawn to scale. The f<strong>or</strong>est tent<br />

caterpillar, Malacosoma disstria (Lepidoptera: Lasiocampidae), (blue with white marks on its back) is a very common<br />

outbreak defoliat<strong>or</strong> of several species of deciduous trees in eastern N<strong>or</strong>th America. The red-headed pine sawfly, Neodiprion<br />

sertifer (Hymenoptera: Diprionidae), is an outbreak species in young f<strong>or</strong>ests of various eastern American pines. The pine<br />

sphinx, Hyloicus pinastri (Lepidoptera: Sphingidae), is a common (green and white-striped) defoliat<strong>or</strong> of pines and spruces in<br />

Europe. The balsam twig aphid, Mindarus sp. (Homoptera: Aphididae), feeds in the spring on balsam fir buds and shoots.<br />

The dead, wilted jack pine shoot, is typical damage caused by the white pine weevil, Pissodes strobi (Coleoptera:<br />

Curculiondiae), a transcontinental species attacking not just pines but also several western spruces in N<strong>or</strong>th America. Gy<strong>or</strong>gi<br />

Csoka provided a photograph of the pine sphinx f<strong>or</strong> artistic inspiration. Kristine A. Kirkeby is the artist.<br />

The use of trade, firm, <strong>or</strong> c<strong>or</strong>p<strong>or</strong>ation names in <strong>this</strong> <strong>publication</strong> is f<strong>or</strong> the inf<strong>or</strong>mation and convenience of the reader.<br />

Such use does not constitute an official end<strong>or</strong>sement <strong>or</strong> approval by the U.S. Department of Agriculture of any<br />

product <strong>or</strong> service to the exclusion of others that may be suitable. Statements of the contribut<strong>or</strong>s from outside the<br />

U.S. Department of Agriculture may not necessarily reflect the policy of the Department.<br />

N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong><br />

F<strong>or</strong>est Service--U.S. Department of Agriculture<br />

1992 Folwell Avenue<br />

St. Paul, Minnesota 55108<br />

Manuscript approved f<strong>or</strong> <strong>publication</strong><br />

1996


DYNAMICS OF FOREST HERBIVORY: QUEST FOR<br />

PATTERN AND PRINCIPLE<br />

Edit<strong>or</strong>s:<br />

William J. Mattson<br />

N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong><br />

Pesticide <strong>Research</strong> Center<br />

Michigan State University<br />

East Lansing, MI 48824 USA<br />

Pekka Niemelii<br />

Faculty of F<strong>or</strong>estry<br />

University of Joensuu<br />

FIN-80101 Joensuu, Finland<br />

Matti Rousi<br />

Punkaharju <strong>Research</strong> <strong>Station</strong><br />

Finnish F<strong>or</strong>est <strong>Research</strong> Institute<br />

FIN-58450 Punkaharju, Finland


PREFACE<br />

This proceedings is the result of an international symposium that was held February 2-6, 1994 in Maui, Hawaii. It was<br />

<strong>or</strong>ganized under the guidelines of the International Union of F<strong>or</strong>estry <strong>Research</strong> Organizations (IUFRO) by the N<strong>or</strong>th<br />

Central F<strong>or</strong>est Experiment <strong>Station</strong> of the U.S. F<strong>or</strong>est Service, and the Finnish F<strong>or</strong>est <strong>Research</strong> Institute. Two IUFRO<br />

W<strong>or</strong>king Parties ($7.01-02, Mechanisms and genetics of plant resistance against insects, and $7.01-03, Mechanisms and<br />

genetics of plant resistance against mammals) convened f<strong>or</strong> <strong>this</strong> international conference on the dynamics of herbiv<strong>or</strong>y<br />

and the elucidation of plant defenses against herbiv<strong>or</strong>es.<br />

The proceedings is published and distributed by the <strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong> in<br />

recognition and unstinting end<strong>or</strong>sement of the IUFRO goals of promoting strong w<strong>or</strong>ld partnerships, particularly in<br />

f<strong>or</strong>est science, f<strong>or</strong> the sustained management, conservation, and the betterment of the W<strong>or</strong>ld's f<strong>or</strong>est and woodland<br />

ecosystems.<br />

ACKNOWLEDGMENTS<br />

The edit<strong>or</strong>s gratefully acknowledge the generous supp<strong>or</strong>t of the N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong> and the Finnish<br />

F<strong>or</strong>est <strong>Research</strong> Institute f<strong>or</strong> underwriting the costs of <strong>or</strong>ganizing, assembling, editing, <strong>print</strong>ing, and distributing <strong>this</strong><br />

<strong>publication</strong>. The edit<strong>or</strong>s also wish to thank all of the particpants whose penetrating and spirited contributions helped<br />

make <strong>this</strong> scientific exchange highly w<strong>or</strong>thwhile.


TABLE OF CONTENTS<br />

Seeking The Rules of Plant-Herbiv<strong>or</strong>e Interactions<br />

l. Why are tree responses to herbiv<strong>or</strong>y so variable?<br />

E. Haukioja and T. Honkanen .............................................................................................................................................. 1<br />

2. Damage-induced nutritional changes in pine foliage: an overview.<br />

R Lyytikainen ..................................................................................................................................................................... I1<br />

3. Stand and landscape diversity as a mechanism of f<strong>or</strong>est resistance to insects.<br />

T. Schowalter ...................................................................................................................................................................... 21<br />

4. On neighb<strong>or</strong> effects in plant-herbiv<strong>or</strong>e interactions.<br />

J. Tuomi, M. Augner, and R Nilsson .................................................................................................................................. 28<br />

5. Defense the<strong>or</strong>ies and birch resistance.<br />

M. Rousi, J. Tahvanainen, W. Mattson, and H. Sikanen .................................................................................................... 39<br />

6. A redox-based mechanism by which environmental stresses elicit changes in plant defensive chemistry.<br />

D.M. N<strong>or</strong>ris ........................................................................................................................................................................ 46<br />

7. Fungal endophytes: Contrasting effects in trees and grasses.<br />

S.H. Faeth ........................................................................................................................................................................... 58<br />

8. Propagating the effects of herbiv<strong>or</strong>e attack in an object-<strong>or</strong>iented model of a tree.<br />

J. Perttunen, H. Saarenmaa, R. Siev_inen, H. Salminen, A. Pouttu, and J. V_ikev_i............................................................ 60<br />

Variable Plants and Herbiv<strong>or</strong>es: Idiosyncrasies of Individual Cases<br />

9. Defoliation of Fagus crenata affects the population dynamics of the beech caterpillar, Quadricalcarifera punctatella.<br />

N. Kamata, Y. Igarashi, and S. Ohara ................................................................................................................................ 68<br />

10. The impact of flowering on the suitability of balsam fir f<strong>or</strong> the spruce budw<strong>or</strong>m varies with larval feeding behavi<strong>or</strong>.<br />

E. Bauce, and N. Carisey .................................................................................................................................................... 86<br />

11. Staminate flowering and tree phenology affect the perf<strong>or</strong>mance of the spruce budw<strong>or</strong>m.<br />

W.J. Mattson, B.A. Birr, and R.K. Lawrence ..................................................................................................................... 97<br />

12. Foliv<strong>or</strong>e feeding on male conifer flowers: defence avoidance <strong>or</strong> bet-hedging?<br />

T.S. Jensen ........................................................................................................................................................................ 104<br />

13. The blackmargined aphid as a keystone species: a predat<strong>or</strong> attract<strong>or</strong> redressing natural enemy<br />

imbalances in pecan systems.<br />

M. K. Harris and T. Li ...................................................................................................................................................... 112<br />

14. Ponderosa pine response to nitrogen fertilization and defoliation by the pand<strong>or</strong>a moth, Col<strong>or</strong>adia pand<strong>or</strong>a Blake<br />

(Lepidoptera: Saturniidae).<br />

B.E. Wickman, R.R. Mason, and H.G. Paul ..................................................................................................................... 118<br />

15. Phytochemical protection and natural enemies: what regulates a root and stem b<strong>or</strong>ing swift moth?<br />

A.V. Whipple, and D.R. Strong ........................................................................................................................................ 127


Genetic Variation in Efficacy of Plant Resistance Against Herbiv<strong>or</strong>es<br />

16. Resistance of hybrid and parental willows to herbiv<strong>or</strong>es: hypotheses and variable herbiv<strong>or</strong>e responses over three years.<br />

R.S. Fritz, B.M. Roche, and C.M. Nichols-Orians ............................................................................................................ 132<br />

17. FI hybrid spruces inherit the phytophagous insects of their parents.<br />

W.J. Mattson, R.K. Lawrence, and B.A. Birr .................................................................................................................... 142<br />

18. Recent advances in research on white pine weevil attacking spruces in British Columbia.<br />

G.K. Kiss, A.D. Yanchuk, R.I. Alfaro, J.E. Carlson, and J.E Manville ........................................................................... 150<br />

19. A sugi clone highly palatable to hares: screening f<strong>or</strong> bioactive chemicals.<br />

H. Hirakawa, T. Nakashima, Y. Hayashi, S. Ohara, and T. Kuwahta .............................................................................. 159<br />

20. Genetic and phenotypic variation in induced reaction of scots pine, Pinus sylvestris L., to Leptographium wingfieMii:<br />

reaction zone length and fungal growth.<br />

F. Lieutier, B. L_ngstr6m, H. Solheim, C. Hellqvist, and A. Yart .................................................................................... 166<br />

2l. Can phloem phenols be used as markers of scots pine resistance to bark beetles.<br />

F. Lieutier, F. Brignolas, V. Picron, A. Yart, and C. Bastiern ............................................................................................ 178<br />

22. Differential susceptibility of white fir provenances to the fir engraver and its fungal symbiont in n<strong>or</strong>thern Calif<strong>or</strong>nia.<br />

G.T. Ferrell, and W.J. Otrosina ......................................................................................................................................... 187<br />

Environmental Variation in Efficacy of Plant Resistance Against Herbiv<strong>or</strong>es<br />

23. Mild drought enhances the resistance of N<strong>or</strong>way spruce to a bark beetle-transmitted blue-stain fungus.<br />

E. Christiansen, and A.M. Glosli ...................................................................................................................................... 192<br />

24. Approaches to studying environmental effects on the resistance of Pinus taeda L. to Dendroctonusfrontalis<br />

Zimmermann.<br />

P. L<strong>or</strong>io ............................................................................................................................................................................. 200<br />

25. Effects of root inhabiting insect-fungal complexes on aspects of tree resistance to bark beetles.<br />

K.F. Raffa, and K. Klepzig ............................................................................................................................................... 211<br />

26. Douglas-fir and western larch defensive reactions to Leptographium abietinum and Ophiostoma pseudotsugae.<br />

D.W. Ross, and H. Solheim ............................................................................................................................................. 224<br />

27. Interruption of bark beetle aggregation by a vig<strong>or</strong>-dependent Pinus host compound.<br />

K.R. Hobson ..................................................................................................................................................................... 228<br />

28. Will global warming alter paper birch susceptibility to bronze birch b<strong>or</strong>er attack?<br />

R.A. Haack....................................................................................................................................................................... 234<br />

29. Variations in spruce needle chemistry and implications f<strong>or</strong> the little spruce sawfly, Pristiph<strong>or</strong>a abietina.<br />

C. Schafellner, R. Berger, J. Mattanovich, and E. Ftihrer .............................................................................................. 248<br />

30. Acidic deposition, drought, and insect herbiv<strong>or</strong>y in an arid environment: Enceliafarinosa and Trirhaba geminata in<br />

southern Calif<strong>or</strong>nia.<br />

T.D. Paine, R.A. Redak, and J.T. Trumble ........................................................................................................................ 257<br />

31. The resistance of scotch pine to defoliat<strong>or</strong>s.<br />

V.I. Grimalsky ................................................................................................................................................................... 263


32. Companion planting of the nitrogen-fixing Gliricidia sepium with the tropical timber species, Milicia excelsa, and its<br />

impact on the gall f<strong>or</strong>ming insect, Phytolyma lata.<br />

M.R. Wagner, J.R. Cobbinah, and D.A. Of<strong>or</strong>i .................................................................................................................. 264<br />

33. Functional heterogeneity of f<strong>or</strong>est landscapes: How host defenses influence epidemiotogy of the southern pine beetle.<br />

R.N. Coulson, J.W. Fitzgerald, B.A. McFadden, RE. Pulley, C.N. Lovelady, and J.R.Gardino ..................................... 272


CONTRIBUTORS and PARTICIPANTS<br />

ALFARO, R.I., Canadian F<strong>or</strong>est Service, Pacific <strong>Research</strong> Centre, Vict<strong>or</strong>ia, B.C., Canada<br />

AUGNER, MAGNUS, Depamnent of Ecology, The<strong>or</strong>etical Ecology, Lund University, T Ecology Bldg., S-22362 Lund,<br />

Sweden<br />

BASTIEN, C., <strong>Station</strong> d'Am61i<strong>or</strong>ation des Arbres F<strong>or</strong>estiers, Institut National de la Recherche Agronomique, Ardon, 45160<br />

Olivet, France<br />

BAUCE, l_,.,D6partement des sciences f<strong>or</strong>esti&es, Facult6 de F<strong>or</strong>esterie et de G6omatique, Universit6 Laval, Ste-Foy<br />

(Qu6bec) G1K 7P4, Canada<br />

BERGER, R., Institute of F<strong>or</strong>est Entomology, F<strong>or</strong>est Pathology, and F<strong>or</strong>est ProtectionUniversit_it ftir Bodenkultur,<br />

Hasenauerstrasse 38, A 1190 Vienna, Austria<br />

BIRR, BRUCE A., <strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong>, B-7 Pesticide <strong>Research</strong> Center, Michigan<br />

State University, East Lansing, M148824, USA<br />

BOWERS, WADE W., Canadian F<strong>or</strong>est Service, Natural Resources Canada, P.O. Box 6028, St. John's, N.F. A 1C 5X8,<br />

Canada<br />

BRIGNOLAS, E, <strong>Station</strong> de Zoologie F<strong>or</strong>esti_re, Institut National de la Recherche Agronomique, Ardon, 45 160 Olivet,<br />

France<br />

CARISEY, W., D6partement des sciences f<strong>or</strong>esti_res, Facult6 de F<strong>or</strong>esterie et de G6omatique, Universit6 Laval, Ste-Foy<br />

(Qu6bec) G1K 7P4, Canada<br />

CARLSON, J.E., University of British Columbia, Biotechnology Lab<strong>or</strong>at<strong>or</strong>y, Vancouver, B.C.<br />

CATES, REX G., Department of Botany, 425WIDB, Direct<strong>or</strong>, Chemical Ecology Lab<strong>or</strong>at<strong>or</strong>y, Brigham Young University,<br />

Provo, UT 84602<br />

CHRISTIANSEN, ERIK, N<strong>or</strong>wegian F<strong>or</strong>est <strong>Research</strong> Institute, N- 1432 As, N<strong>or</strong>way<br />

COBBINAH, JOSEPH R., F<strong>or</strong>estry <strong>Research</strong> Institute of Ghana, UST P.O. 63, Kumasi, Ghana<br />

COULSON, ROBERT N., Knowledge Engineering Lab<strong>or</strong>at<strong>or</strong>y, Departments of Entomology and Geography, Texas A&M<br />

University, College <strong>Station</strong>, TX 77843<br />

FAETH, STANLEY H., Department of Zoology, Arizona State University, Tempe, Arizona 85287-1501, USA<br />

FERRELL, GEORGE T., <strong>USDA</strong> F<strong>or</strong>est Service, Pacific Southwest <strong>Research</strong> <strong>Station</strong>, 2600 Washington Avenue, Redding,<br />

Calif<strong>or</strong>nia 96001, USA<br />

FITZGERALD, JEFFREY W., Knowledge Engineering Lab<strong>or</strong>at<strong>or</strong>y, Departments of Entomology and Geography, Texas<br />

A&M University, College <strong>Station</strong>, TX 77843<br />

FRITZ, ROBERT S., Department of Biology, Box 133, Vassar College, Poughkeepsie, NY 12601, USA<br />

F_)HRER, E., Institute of F<strong>or</strong>est Entomology, F<strong>or</strong>est Pathology, and F<strong>or</strong>est Protection, Universit_it ftir Bodenkultur,<br />

Hasenauerstrasse 38, A 1190 Vienna, Austria<br />

GIARDINO, JOHN R., Knowledge Engineering Lab<strong>or</strong>at<strong>or</strong>y, Departments of Entomology and Geography, Texas A&M<br />

University, College <strong>Station</strong>, TX 77843<br />

GLOSLI, ANNE MARIE, N<strong>or</strong>wegian F<strong>or</strong>est <strong>Research</strong> Institute, N- 1432 As, N<strong>or</strong>way


GRIMALSKY, ¥.I., <strong>Research</strong> Institute of F<strong>or</strong>estry, Gemol, Bel<strong>or</strong>ussia<br />

HAACK, ROBERT A., <strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong>, 1407 S. Harrison Road, East<br />

Lansing, MI 48823, USA<br />

HARRIS, MARVIN K., Department of Entomology, Texas A&M University, College <strong>Station</strong>, TX 77843<br />

HAUKIOJA, ERKKI, Lab<strong>or</strong>at<strong>or</strong>y of Ecological Zoology, Department of Biology, University of Turku, FIN-20500 Turku,<br />

Finland<br />

HAYASHI, YOSHIOKI, F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute, PO Box 16, Tsukuba-N<strong>or</strong>in, Ibaraki 305, Japan<br />

HELLQVIST, C., Division of F<strong>or</strong>est Entomology, Swedish University of Agricultural Sciences, S-77073 Garpenberg,<br />

Sweden<br />

tlELSON, BLAIR, F<strong>or</strong>est Pest Management Institute, Canadian F<strong>or</strong>est Service, Natural Resources Canada, Sault Ste. Marie,<br />

Ontario, P6A 4J, Canada.<br />

HERMS, DANIEL A., The DOW Gardens, 1018 W. Main Street, Midland, MI 48640<br />

HIRAKAWA, HIROFUMI, F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute, PO Box 16, Tsukuba-N<strong>or</strong>in, Ibaraki 305, Japan<br />

HOBSON, KENNETH R., Department of Entomology, 345 Russell Labs, University of Wisconsin, Madison, WI 53706,<br />

USA<br />

HONKANEN, TUI,IA, Lab<strong>or</strong>at<strong>or</strong>y of Ecological Zoology, Department of Biology, University of Turku, FIN-20500 Turku,<br />

Finland<br />

IGARASHI, YUTAKA, Lab<strong>or</strong>at<strong>or</strong>y of F<strong>or</strong>est Entomology, Tohoku <strong>Research</strong> Center, F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong><br />

Institute, Nabeyashiki 72, Shimokuriyagawa, M<strong>or</strong>ioka, Iwate 020-01, Japan<br />

ISAEV, ALEX, International F<strong>or</strong>estry Institute, Udalcov Str. 24.87, Moscow, Russia<br />

JENSEN, THOMAS SECHER, Institute of Biological Sciences, Department of Zoology, Aarhus University, DK-8000<br />

Aarhus C, Denmark<br />

KAMATA, NAOTO, Lab<strong>or</strong>at<strong>or</strong>y of F<strong>or</strong>est Entomology, Tohoku <strong>Research</strong> Center, F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong><br />

Institute, Nabeyashiki 72, Shimokuriyagawa, M<strong>or</strong>ioka, Iwate 020-01, Japan<br />

KISS, GYULA K., B.C. F<strong>or</strong>est Service, Kalamalka F<strong>or</strong>estry Centre, 3401 Reservoir Road, Vernon, B.C. V1B 2C7<br />

KLEPZIG, KIER D., Department of Plant & Soil Science, Southern University and A&M College, Baton Rouge, LA<br />

70813, USA<br />

KUMAPLEY, PHILOMENA A., Institute of Renewable Natural Resources, University of Science & Technology, Kumasi-<br />

Ghana, W. Africa<br />

KUWAItATA, TSUTOMU, F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute, PO Box 16, Tsukuba-N<strong>or</strong>in, Ibaraki 305, Japan<br />

LANGSTRt_M, B., Department of Plant & F<strong>or</strong>est Protection, Swedish Univ. of Agric. Science, Box 7044, S-75007 Uppsala,<br />

Sweden<br />

LARSSON, STIG, Department of Plant & F<strong>or</strong>est Protection, Swedish Univ. of Agric. Science, Box 7044, S-75007 Uppsala,<br />

Sweden<br />

LAWRENCE, ROBERT K., <strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong>, B-7 Pesticide <strong>Research</strong><br />

Center, Michigan State University, East Lansing, MI 48824, USA


LI, T., Department of Entomology, Texas A&M University, College <strong>Station</strong>, TX 77843<br />

LIEUTIER, E, <strong>Station</strong> de Zoologic F<strong>or</strong>esti?_re, Institut National de la Recherche Agronomique, Ardon, 45160 Olivet, France<br />

LORIO, PETER L., Jr., <strong>USDA</strong> F<strong>or</strong>est Service, Southern F<strong>or</strong>est Experiment <strong>Station</strong>, 2500 Shrevep<strong>or</strong>t Highway, Pineville,<br />

LA 71360, USA<br />

LOVELADY, CLARK N., Knowledge Engineering Lab<strong>or</strong>at<strong>or</strong>y, Departments of Entomology and Geography, Texas A&M<br />

University, College <strong>Station</strong>, TX 77843<br />

LIEUTIER, F., <strong>Station</strong> de Zoo[ogie F<strong>or</strong>esti6re, Institut National de la Recherche Agronomique, Ardon, 45160 Olivet, France<br />

LYYTIK,g, INEN, P)/,IVI, Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Department of F<strong>or</strong>est Ecology, EO. Box 18, FIN-01301 Vantaa,<br />

Finland<br />

MANVILLE, J.E, Canadian F<strong>or</strong>est Service, Pacific <strong>Research</strong> Centre, Vict<strong>or</strong>ia, B.C.<br />

MASON, R.R., F<strong>or</strong>estry and Range Sciences Lab<strong>or</strong>at<strong>or</strong>y, 1401 Gekeler Lane, La Grande, OR 97850<br />

MATTANOVICH, J., Institute of F<strong>or</strong>est Entomology, F<strong>or</strong>est Pathology, and F<strong>or</strong>est ProtectionUniversit_it fi.irBodenkultur,<br />

Hasenauerstrasse 38, A 1190 Vienna, Austria<br />

MATTSON, WILLIAM J., <strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong>, B-7 Pesticide <strong>Research</strong> Center,<br />

Michigan State University, East Lansing, MI 48824, USA<br />

McFADDEN, BRYAN A., Knowledge Engineering Lab<strong>or</strong>at<strong>or</strong>y, Departments of Entomology and Geography, Texas A&M<br />

University, College <strong>Station</strong>, TX 77843<br />

NAKASHIMA, TADAKAZU, F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute, PO Box 16,Tsukuba-N<strong>or</strong>in, Ibaraki 305,<br />

Japan<br />

NIEMEL,_, PEKKA, Faculty of F<strong>or</strong>estry, University of Joensuu, FIN-80101, Joensuu, Finland<br />

NILSSON, PATRIC, Department of Ecology, The<strong>or</strong>etical Ecology, Lund University, T Ecology Bldg., S-22362 Lund,<br />

Sweden<br />

NORRIS, DALE M., Department of Entomology, 642 Russell Lab<strong>or</strong>at<strong>or</strong>ies, University of Wisconsin, Madison, WI 53706,<br />

USA<br />

OFORI, DANIEL A., F<strong>or</strong>estry <strong>Research</strong> Institute of Ghana, UST EO. 63, Kumasi, Ghana<br />

OHARA, SEIJI, Lab<strong>or</strong>at<strong>or</strong>y of Chemical Conversion, Wood Chemistry Division, F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong><br />

Institute, Matshunosato 1, Kukizaki, Inashiki, Ibaraki 305, Japan<br />

ORIANS, COLIN M., Department of Biology, Williams College, Williamstown, MA 01267, USA<br />

OTROSINA, W.J., <strong>USDA</strong> F<strong>or</strong>est Service, Southeastern F<strong>or</strong>est Experiment <strong>Station</strong>, 320 Green Street, Athens, Ge<strong>or</strong>gia<br />

30602, USA<br />

PAINE, 'I.D., Department of Entomology, University of Calif<strong>or</strong>nia, Riverside, CA 92521<br />

PAUL, H.G., F<strong>or</strong>estry and Range Sciences Lab<strong>or</strong>at<strong>or</strong>y, 1401 Gekeler Lane, La Grande, OR 97850<br />

PERTTUNEN, JARI, Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Unioninkatu 40A, SF-00170, Helsinki, Finland<br />

PICRON, V., <strong>Station</strong> de Zoologie F<strong>or</strong>esti_re, Institut National de la Recherche Agronomique, Ardon, 45160 Olivet, France


POUTTU, ANTTI, Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Unioninkatu 40A, SF-00170, Helsinki, Finland<br />

PULLEY, PAUL E.. Knowledge Engineering Lab<strong>or</strong>at<strong>or</strong>y, Departments of Entomology and Geography, Texas A&M UI_iversity,<br />

College <strong>Station</strong>, TX 77843<br />

QUIRING, DANIEL T., Pop. Ecol. Group, Department of F<strong>or</strong>est Resources, Univeristy of New Brunswick. Fredericton,<br />

N.B. E3B6C2<br />

RAFFA, KENNETH f'., Department of' Entomology, 642 Russell Lab<strong>or</strong>at<strong>or</strong>ies, University of Wisconsin, Madison, Wt<br />

53706. USA<br />

REDAK, R.A., Depamnent of Entomology, University of Calif<strong>or</strong>nia, Riverside, CA 92521<br />

ROCHE, BERNADETTE M., Department of Biology, Vassar College, Poughkeepsie, NY 12601, USA<br />

ROSS, DARRELL W., Department of F<strong>or</strong>est Science, Oregon State University, C<strong>or</strong>vallis, Oregon 97331-7501. USA<br />

ROUSI, MATTI, The Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Punkaharju <strong>Research</strong> <strong>Station</strong>, SF-58450 Punkaharju, Finland<br />

SAARENMAA, ttANNU, Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Unioninkatu 40A, SF-00170, Helsinki, Finland<br />

SALMINEN, HANNU, Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Unioninkatu 40A, SF-00170, Helsinki, Finland<br />

SCHAFELI,NER, CHRISTA, Institute of F<strong>or</strong>est Entomology, F<strong>or</strong>est Pathology, and F<strong>or</strong>est Protection, Universit;:it f/Jr<br />

Bodenkultur, Hasenauerstrasse 38, A 1190 Vienna, Austria<br />

SCHOWAI2FER, TIM D,, Department of Entomology, Oregon State University, C<strong>or</strong>vallis, Oregon 97331-2907, USA<br />

SIEVANEN, R|STO, Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Unioninkatu 40A, SF-(X) IT0, Helsinki, Finland<br />

SIKANEN, ttANNI, The Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Punkaharju <strong>Research</strong> <strong>Station</strong>, SF-58450 Punkaharju, Finland<br />

SOLHEIM, HAI,VOR, N<strong>or</strong>wegian F<strong>or</strong>est <strong>Research</strong> Institute, Section of F<strong>or</strong>est Ecology, Division of F<strong>or</strong>est Pathology, N-<br />

1423 As-NLH, N<strong>or</strong>way<br />

STRONG, DONALD R., Bodega Marine Lab<strong>or</strong>at<strong>or</strong>y, University of Calif<strong>or</strong>nia, Box 247, Bodega Bay, CA 94923<br />

TAttVANAINEN, JORMA,University of Joensuu, Department of Biology, P.O. Box 111, SF-80101 Joensuu, Finland<br />

TRUMBLE, J,T., Department of Entomology, University of Calif<strong>or</strong>nia, Riverside, CA 92521<br />

TUOMI, JUHA, Department of Ecology, The<strong>or</strong>etical Ecology, Lund University, T Ecology Bldg., S-223 62 Lund, Sweden<br />

VAKEVA, JOUNI, Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Unioninkatu 40A, SF-00170, I-telsinki, Finland<br />

WAGNER, MICIIAEL R., School of F<strong>or</strong>estry, N<strong>or</strong>thern Arizona University, RO. Box 15018, Flagstaff, AZ 86011, USA<br />

WHIPPLE, AMY V., Bodega Marine Lab<strong>or</strong>at<strong>or</strong>y, University of Calif<strong>or</strong>nia, Box 247, Bodega Bay, CA 94923<br />

WICKMAN, B.E., <strong>USDA</strong> F<strong>or</strong>est Service, Silviculture Lab<strong>or</strong>at<strong>or</strong>y, 1027 NW Trenton Avenue, Bend, OR 97701<br />

YANCHUK, A.D., B.C. F<strong>or</strong>est Service, <strong>Research</strong> Branch, 31 Bastion Square, Vict<strong>or</strong>ia, B.C. V8W 3E7E<br />

YART, A., <strong>Station</strong> de Zootogie F<strong>or</strong>estibre, Institut National de la Recherche Agronomique, Ardon, 45160 Olivet, France<br />

YOSHIKAWA, KEN, Fac. of Agriculture, Okayama Univ., Tsushima-naka 1- l- 1, Okayama 700, Japan


WHY ARE TREE RESPONSES TO HERBIVORY SO VARIABLE?<br />

ERKKI HAUKIOJA and TUIJA HONKANEN<br />

Lab<strong>or</strong>at<strong>or</strong>y of Ecological Zoology, t)epartment of Biology, University of Turku<br />

FIN-20500 Turku, Finland<br />

INTRODUCTION<br />

Following defoliation the quality of tree leaves can become 'w<strong>or</strong>se, better, <strong>or</strong> remain unchanged f<strong>or</strong> herbiv<strong>or</strong>es. To some<br />

extent, such variation results from tile different types of foliage which are studied. F<strong>or</strong> example, after severe defoliation,<br />

trees may reflush in the same season, but these new leaves differ from n<strong>or</strong>mal, mature foliage, and leaves which grew at the<br />

time of partial defoliation may become different fi'om leaves in an undamaged tree. Similarly, defoliated and undefoliated<br />

trees may produce qualitatively different foliage in the next growth season. The fk)liage produced in the year after defoliation<br />

occurred has been found to be especially different from that of undefoliated trees both in chemical traits, and herbiv<strong>or</strong>e<br />

bioassays, as is the case f<strong>or</strong> birch (Neuvonen and Haukioja 1991), larch (Benz 1974), and oak (Rossiter et al. 1988). All of<br />

_:hese trees are deciduous and there seems to be a general trend that defoliation alters the quality of deciduous foliage m<strong>or</strong>e<br />

noticeably than that of evergreen f_liage.<br />

We have studiec/defoliation-induced responses in mountain birch, Betula pubescens ssp t<strong>or</strong>tuosa, and Scots pine, Pinus<br />

s)'Ivestris, i.e., in a deciduous species responding strongly to herbiv<strong>or</strong>y (Haukioja et al. 1985), and in an evergreen species<br />

tending toward the other end of the continuum (Niemelfi et al. 1984, 1991, Watt et al. 1991). In <strong>this</strong> paper, we briefly review<br />

the treatments which are beneficial (induced ameli<strong>or</strong>ation, IA hereafter) <strong>or</strong> harmful (induced resistance, IR hereafter) to<br />

herbiv<strong>or</strong>es. Furtherm<strong>or</strong>e, we discuss the validity of several hypotheses which attempt to explain how and why trees respond<br />

to herbiv<strong>or</strong>y as they do: the carbon-nutrient (C/N) balance hypothesis (Bryant et al. 1983, 1991; Tuomi et al. 1991), the<br />

growth--differentiation (G/D) balance hypothesis (Herrns and Mattson 1992), and the sink-source (S/S) hypothesis (Haukioja<br />

1991b, Honkanen et al. 1994, Honkanen and ttaukioja 1994).<br />

TIlE TARGET TISSUE AND TIMING OF THE DAMAGE<br />

Whether IA <strong>or</strong> IR is elicited after damage does not depend upon the species <strong>or</strong> individual concerned. The same tree<br />

may become either better <strong>or</strong> w<strong>or</strong>se depending on the type of tissue which was damaged (Haukioja et al. 1990), and also the<br />

timing of damage (Neuvonen et al. 1988).<br />

IR can be separated into rapid (RIR) and delayed (DIR) induced resistance (Haukioja 1982). In the f<strong>or</strong>mer, po<strong>or</strong> diet<br />

quality affects the generation of herbiv<strong>or</strong>es which caused the damage, while the latter influences the next generation(s) of<br />

herbiv<strong>or</strong>es. Note that <strong>this</strong> division is relevant as regards the consequences of the damage on herbiv<strong>or</strong>e populations dynamics:<br />

RIR is a stabilizing and DIR a destabilizing fact<strong>or</strong> because it introduces time lags. Furtherrn<strong>or</strong>e, the division into RIR and<br />

DIR is meaningful only in the context of a herbiv<strong>or</strong>e generation time; a certain response by the plant may be experienced as<br />

DIR by a sh<strong>or</strong>t-lived herbiv<strong>or</strong>e species but as RIR by a long-lived one. From the plant's point of view these rapid and<br />

delayed responses do not necessarily differ, A c<strong>or</strong>responding distinction between rapid and delayed responses also applies to<br />

induced ameli<strong>or</strong>ation.<br />

The mountain birch system provides an example of how herbiv<strong>or</strong>e-induced damage results in a multitude of changes<br />

in foliage quality. Po<strong>or</strong> quality leaves are produced after laminae of growing <strong>or</strong> newly expanded leaves are damaged by real<br />

<strong>or</strong> simulated herbiv<strong>or</strong>y. "['he RIR in birches tends to be mild when measured by its effects on herbiv<strong>or</strong>es (Haukioja and<br />

Neuvonen 1987), while DIR may severely curtail the potential increase of the herbiv<strong>or</strong>e population (Haukioja et al. 1985).<br />

Attempts to trigger RIR <strong>or</strong> DIR by damaging mature, late summer foliage have led to mild <strong>or</strong> no responses at all (Neuvonen<br />

et al. 1988, Tuomi et al. 1988b, Hanhim_iki and Senn 1992). On the other hand, the effects of early season damage may<br />

Mattson, W.J., Niemel_,1, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.


persist in mature leaves and affect subsequent herbiv<strong>or</strong>es, even from different guilds (Neuvonen et al. 1988, HanhimS.ki<br />

1989). The delayed effects (DIR) can remain effective on foliv<strong>or</strong>es consuming mature leaves produced the following<br />

summer (Hanhimfiki 1989). Induced resistance in mountain birch foliage is triggered most effectively by damage at the time<br />

when there is peak feeding by the most destructive insect pest of mountain birch, the geometrid, Epirrita autumnata (Tenow<br />

1972, Haukioja et al. 1988). Since defoliation tends to effect all members of the birch chewing guild similarly (Hanhim_iki<br />

1989), the outcomes of defoliation in birch foliage seem to represent a general response, and are not specifically targeted<br />

against Epirrita autumnata <strong>or</strong> other early season defoliat<strong>or</strong>s.<br />

Induced ameli<strong>or</strong>ation in mountain birch is triggered either by removal of the apical tips of d<strong>or</strong>mant twigs (Haukioja et<br />

al. 1990) <strong>or</strong> by damage to expanding buds <strong>or</strong> tiny leaves (Senn and Haukioja 1994). In practice, the f<strong>or</strong>mer type of damage<br />

may result from winter browsing by mammals and the latter from insect damage very early in the season. Winter browsing<br />

makes the f<strong>or</strong>thcoming foliage m<strong>or</strong>e preferable f<strong>or</strong> a suite of different types of herbiv<strong>or</strong>es (Danell and Huss-Danell 1985),<br />

although the effects on insect perf<strong>or</strong>mance tend to be mild (Neuvonen and Danell 1987, Haukioja et al. 1990). Larvae of<br />

Epirrita autumnata start feeding at, <strong>or</strong> even bef<strong>or</strong>e, budbreak, and are known to destroy apical buds from branches (Haukioja<br />

et al. 1990). Early leaf damage might affect foliage quality in the same year while bud removal influences growth in the<br />

following season. In practice <strong>this</strong> means that depending on the type of target tissue and timing of the damage, larvae of<br />

Epirrita autumnata are potentially capable of inducing either IA, RIR <strong>or</strong> DIR in birch foliage (Haukioja 1991a).<br />

The case with the pine is different. Most studies conducted on pines have not rep<strong>or</strong>ted strong qualitative changes in<br />

needle chemistry after defoliation and, acc<strong>or</strong>dingly, the effects on insect perf<strong>or</strong>mance also have been mild <strong>or</strong> insignificant<br />

(e.g., NiemelS. et al. 1991, Watt et al. 1991, Reich et al. 1993). However, some studies have revealed defoliation-induced<br />

chemical changes in pines. Wagner and Evans (1985) rep<strong>or</strong>ted that defoliation of Pinus ponderosa seedlings increased the<br />

production of both phenols and proteins in mature and immature foliage. Honkanen and Haukioja (1994) demonstrated<br />

increased nitrogen needle concentration after severe defoliation of Scots pine. These results indicate that the outcomes of<br />

defoliation in pines can vary substantially, and so are the consequences on herbiv<strong>or</strong>es.<br />

Pruning <strong>or</strong> removal of pine buds <strong>or</strong> twigs causes m<strong>or</strong>e lush growth of needles and shoots (e.g., Nu<strong>or</strong>teva and Kurkela<br />

1993, Honkanen et al. 1994), just like in mountain birch. Pine browsers, particularly moose, are known to cause IA<br />

(L6yttyniemi 1985), but it is not known whether insect herbiv<strong>or</strong>es of pine are able to induce such a response.<br />

In summary, both in the case of mountain birch and Scots pine it is possible to induce variable effects - better <strong>or</strong><br />

w<strong>or</strong>se foliage quality, <strong>or</strong> no change at all - by selectively damaging different types of tissues and by varying the timing of the<br />

treatment.<br />

THE RESOURCE STATUS OF BIRCH AND PINE FOR INDUCED RESPONSES:<br />

ADEQUACY OF THE C/N HYPOTHESIS<br />

The simplest explanation f<strong>or</strong> the induced resistance is that chewing <strong>or</strong> browsing of plant biomass removes nutrients<br />

and such losses lower foliage quality f<strong>or</strong> later herbiv<strong>or</strong>es (Tuomi et al. 1984). If the loss of nutrients is critical f<strong>or</strong> induced<br />

resistance, the tree's status can be remedied by fertilization which usually increases the nutrient concentration in foliage. The<br />

crucial question here is whether fertilization decreases the resistance of the tree and cancels any negative effects of defoliation.<br />

In the case of the mountain birch, Haukioja and Neuvonen (1985) rep<strong>or</strong>ted that nitrogen fertilization did not alleviate<br />

the detrimental effects of defoliation (DIR) on Epirrita larvae. This contradicts the predictions of the carbon-nutrient balance<br />

(C/N) hypothesis, which ascribes the resistance of deciduous trees to carbon-based secondary compounds which generally<br />

decrease when trees obtain extra nutrients. However, Bryant et al. (1993) found that fertilized paper birches, Betula<br />

papyrifera, lost their ability f<strong>or</strong> DIR. They criticized the mountain birch experiment by Haukioja and Neuvonen (1985) f<strong>or</strong><br />

simultaneous application of the fertilizing and defoliation treatments; the trees might not have had time to benefit from<br />

increased nutrient availability. However, another experiment with the mountain birch-Epirrita system (Ruohom_iki et al. in<br />

press), accounted f<strong>or</strong> the criticism by Bryant et al. (1993) and yet rep<strong>or</strong>ted the same result of Haukioja and Neuvonen (1985).<br />

Furtherm<strong>or</strong>e, the strength of the DIR did not decline with shading, as predicted by the C/N hypothesis (Ruohom_iki et al. in<br />

press). It is possible that mountain birch and paper birch behave differently, <strong>or</strong> the insects used in the two experiments<br />

showed differential tolerance to the altered carbon-based compounds, <strong>or</strong> that changes in carbon and nitrogen were not crucial<br />

f<strong>or</strong> the herbiv<strong>or</strong>e. In any case, it is obvious that f<strong>or</strong> birches the C/N hypothesis did not provide the general explanation f<strong>or</strong><br />

induced responses.<br />

2


The C/N hypothesis does not adequately predict fertilization induced changes in pine foliage, either. Honkanen et al.<br />

(1994) and H<strong>or</strong>_kanen and Hauki_ja (1994) demonstrated that branches in fertilized Scots pi_es produced m<strong>or</strong>e and bigger<br />

needles irrespective of whether the branch <strong>or</strong> the whole tree was del'\)liated <strong>or</strong> not. This is hard to reconcile with tile predic-<br />

tion that carbon deficiency after defoliation restricts growth of foliage in evergreens (e.g., Bryant et al. I983, Tuomi et al.<br />

19882). Furtherm<strong>or</strong>e, fertilization does not generally decrease, and may even increase, allocation to carbon-based secondary<br />

compounds in Scots pine (Bj/.Srkman et al. 199t, Edenius 1993).<br />

in s_._mmary, the C/N hypothesis does not adequately explain the effects of changes in the availability of mineral<br />

nutrients <strong>or</strong> shade f<strong>or</strong> DIR in birch, <strong>or</strong> the f<strong>or</strong>mer in pine. Sit-_ce the growth-differentiation balance (G/D) hypothesis is based<br />

on the C/N hypothesis with respect to resource availability, we doubt the generality of both these to explain herbiv<strong>or</strong>e-<br />

induced responses of trees on the basis of resource availability (Iason and Hest,er 1993).<br />

THE FUNCTIONAL ORGANIZATION OF THE TREE<br />

Next we briefly discuss the functional <strong>or</strong>ga_aization of trees to demonstrate that some very basic physiological<br />

mechanisms behind tree structure and function actually contribute to a simple, ahnost mechanistic explanation f<strong>or</strong> induced<br />

responses of: birch. The same basic explanation may also contribute to understanding the differences between responses of<br />

the deciduous mountain birch and the evergreen Scots pine.. The tkfllowing reasoning f<strong>or</strong>ms the basis of tile source-sink (S/S)<br />

hypothesis. First, trees are modular <strong>or</strong>ganisms. Usually <strong>this</strong> statement is interpreted as a description of the m<strong>or</strong>phological<br />

desigrl of trees: they are built of repetitive, multicellular units comprising a stem segment, leaf and meristem (White 19791).<br />

This emphasis does not help much in explaining tree functions although it vaguely makes understandable that tree parts may<br />

respond semioindependently, contrary to parts of unitary <strong>or</strong>ganisms like humans <strong>or</strong> fruit flies (Vu<strong>or</strong>isalo and Tuomi 1986).<br />

The term integrated physiological unit (IPU) was coined f<strong>or</strong> <strong>this</strong> purpose (Watson and Casper 1984, Watson 1986). That<br />

concept emphasizes m<strong>or</strong>e the independence of IPUs than the integration among IPU's. However, it is the latter that makes a<br />

tree function like one individual. Haukioja (1991b) proposed that a simple set of rules can generally determine the integrity<br />

of tree functions and help not only to understand the relative independence of IPU's but also the temp<strong>or</strong>al and spatial variabil-<br />

ity in their independence. Furtherm<strong>or</strong>e, it explicitly acknowledges other parts of the tree besides the canopy, i.e., the stem<br />

and roots.<br />

Integration within an individual tree simply requires that IPUs at the "best" locations gain control over other, less<br />

optimally located IPUs. Within the canopy, <strong>this</strong> happens if the apical rneristems, which have the best access to light, remain<br />

<strong>or</strong> become metabolically active and simultaneously suppress nearby, less tav<strong>or</strong>ably located, meristems (Haukioja 1991b,<br />

Sachs e,ral. 1993). F<strong>or</strong> <strong>this</strong> outcome, resources should be preferentially transp<strong>or</strong>ted to the winning meristems. This takes<br />

place by their elevated activity relative to that of nearby meristems; high meristematic activity is coupled with the production<br />

of pertinent growth h<strong>or</strong>mones which, in turn, preferentially direct the flow of resources to these strong meristems. We<br />

contend that <strong>this</strong> incidentally creates the integration among above-ground parts of the plant. Integration within a tree also<br />

varies seasonally (Sprugel et al. 199t). We assume <strong>this</strong> results from the existence of other meristems than apical ones in the<br />

canopy. Trees have lateral meristems in branches, trunks and roots. When the sink strength of these exceeds that of apical<br />

meristems, resources are preferentially conveyed to the trunk and to the below-ground parts. When <strong>this</strong> happens, the whole<br />

tree is well integrated, and the m<strong>or</strong>phologically defined IPU's have temp<strong>or</strong>ally lost their functional semi-independence.<br />

The responses of IPU's ultimately are based on genetic rules which are facultative and subject to position-dependent<br />

influences. Thus, irrespective of the similar genetic basis, the behavi<strong>or</strong> of an IPU depends on its position within the plant<br />

(which takes into account the possible dominance of other, especially nearby, IPUs) and on the external environment. The<br />

latter includes availability of resources. Theref<strong>or</strong>e, the potential and the actual behavi<strong>or</strong> of an IPU may be constrained by<br />

either internal (genetic and position-dependent) <strong>or</strong> external (resources, competition) fact<strong>or</strong>s, <strong>or</strong> by both. The production of a<br />

tree-like architecture requires that the genetically determined facultative and position-dependent rules include instructions f<strong>or</strong><br />

achieving, maintaining and relinquishing dominance among competing IPUs. However, we assume that, within the limits of<br />

the genetically based conditional and position-dependent rules, competition among IPUs within an individual plant is<br />

physiologically genuine yet these rules still transtk_rm the population of IPUs into an integrated "individual plant".<br />

Variance in the strength of apical dominance (in addition to traits like d<strong>or</strong>mancy/activity of n_teristems, internode<br />

length and branching angle, Suthertand and Stillman 1990) explains a lot of differences among different tree types. Since the<br />

relative strength of apical and other meristems has a hereditary basis (Kozlowski 1971), characteristics like the shape of the<br />

tree (resulting from the intermodular interactions) are susceptible to artificial and natural selection.<br />

3


THE SINK-SOURCE HYPOTHESIS: TYPE AND TIMING OF DAMAGE ARE CRUCIAL<br />

A crucial point underlying variable responses after damage to different types of tissues is that different meristems are<br />

provisioned seasonally at different times (Fig. 1). We claim that such a predictable temp<strong>or</strong>al variation in allocation of plant<br />

resources to different sinks, as well as shifts of individual meristems from sinks to sources within the course of the season,<br />

provides an imp<strong>or</strong>tant explanation f<strong>or</strong> the variability in damage-induced responses of birch and pine foliage (Table 1).<br />

MOUNTAIN BIRCH<br />

Early spring sink leaves < _-- resource pool<br />

Early summer young source leaves -> local sinks<br />

Late summer mature source leaves > resource pool<br />

SCOTS PINE<br />

Early spring source leaves ---- _ resource pool<br />

Early/late summer apical source leaves _ local sinks<br />

basal source leaves-- > resource pool<br />

Figure l.--The assumed seasonal variation in resource flows in the mountain birch and Scots pine.<br />

Flushing leaves are strong sinks, and we assume that their success in drawing resources depends on their sink<br />

strength. Within few days young leaves shift from sinks to sources (Valanne and Valanne 1984) and start to preferentially<br />

provision nearby buds which will contain the leaf prim<strong>or</strong>dia f<strong>or</strong> the next season. Simultaneously the leaves themselves<br />

mature, and when the growth of buds in sh<strong>or</strong>t shoots, and of meristem-supp<strong>or</strong>ting leaves has been completed, we assume<br />

lateral meristems and meristems in the root system to be the dominant sinks tbr resources.<br />

Damage to Physiological Sinks in Birch<br />

Consequences of damage to sink <strong>or</strong>gans depend on the strength and the position of the sinks. Sinks can be eliminated<br />

<strong>or</strong> weakened by clipping the tips of d<strong>or</strong>mant shoots, by removing the apical d<strong>or</strong>mant buds only (Haukioja et al. 1990), <strong>or</strong>,<br />

later in spring, by damaging flushing apical leaves which still are in the strong sink phase (Senn and Haukioja 1994). This<br />

releases extra resources f<strong>or</strong> n<strong>or</strong>mally suppressed basal meristems (Lehtil_ et al. ms.). The key is the apical dominance and<br />

not the amount of nutrients lost in the damage because when small but equal amounts of basal, instead of apical, meristems<br />

were removed, no changes were found in the extant foliage (Haukioja et al. 1990).<br />

Damage to Physiological Sources in Birch<br />

The consequences of damage to birch source leaves depend on the meristems which the leaves were feeding at the<br />

time of the damage (and which they would have provisioned had the damage not taken place). Damage to young source<br />

leaves reduces the sink strength of meristems in the leaf and stem prim<strong>or</strong>dia which the relevant leaves were feeding at the<br />

moment of the damage. F<strong>or</strong> birch long-shoot leaves, those meristems are located in their axillary buds and f<strong>or</strong> sh<strong>or</strong>z-shoot<br />

leaves (which f<strong>or</strong>m the vast maj<strong>or</strong>ity of all leaves in mature mountain birches) they are in sh<strong>or</strong>t shoot buds from which the<br />

next season leaves flush. Both suffer significantly from even very limited local damage; the shoot from the axillary bud a of<br />

damaged long shoot leaf grows sh<strong>or</strong>t the next year (Haukioja et al. 1990, Ruohom_iki et al. ms). Similarly, leaves as well as<br />

4


Table l.---Typical changes in %liage growth vig<strong>or</strong>, nitrogen concentration and the amount of secondary compounds after<br />

different types of damage. + = increase, 0 = no change and - = decrease in the studied character.<br />

Damaged tissue Growth Nitrogen Secondary Reference<br />

vig<strong>or</strong> compounds<br />

BIRCH<br />

Buds/flushing leaves + + 1,2<br />

Young source leaves - - + 2,3,4,5<br />

Mature source leaves 0 0 0 5,6<br />

SCOTS PINE<br />

Buds/flushing leaves + + 9 7,8<br />

Source needles<br />

feeding current<br />

year foliage 0,+ ;' 0 7,9,10,11<br />

Source needles<br />

feeding general<br />

st<strong>or</strong>es 0 0 0 9,10<br />

l) Danell et al. (unpblished data); 2) Niemelfi er al. 11979;3) Haukioja er al. 1985; 4) Haukioja er al. 1990; 5) Tuomi er al.<br />

1988; 6) Hartley and Lawton 1991; 7) Honkanen et al. 1994; 8)Lgmgstr6m e/al. 1990; 9) Ericsson er al. 1980; 10) Honkanen<br />

and Haukioja ( 1994); 11) Niemel/i et al. 1991<br />

_Depends on the scale of defbliation (Honkanen and Haukioja 1994)<br />

catkins in sh<strong>or</strong>t shoots, whose leaves were damaged the previous year, remain smaller than in undamaged shoots (Ruohom_iki<br />

egal. ms.). Since these outcomes follow from damage to even a single leaf, reduction of the total resource pool of the tree<br />

cannot be the relevant explanation in such a case. Instead, the meristems fed by damaged leaves presumably were unable to<br />

efficiently compete tot resources in the common pool (Ruohom_iki er at. ms).<br />

Damage to late season foliage does not affect the sink strength of the following year's buds because they had been<br />

completed earlier. Instead, <strong>this</strong> may :reduce st<strong>or</strong>ed resources.<br />

The above scenario offers the simplest explanation f<strong>or</strong> DIR: early summer defoliation during the previous growth<br />

season led to "weak" buds which, when still in the sink phase, were unable to get their n<strong>or</strong>mal load of nutrients and photosynthates.<br />

The po<strong>or</strong> quality of such foliage f<strong>or</strong> herbiv<strong>or</strong>es may follow both from reduced amounts of nutrients and from increased<br />

secondary compounds. The same basic explanation might apply to local RIR, too: when foliage is damaged in a<br />

branch, the developing foliage loses part of its sink strength, and the whole branch might theref<strong>or</strong>e become a po<strong>or</strong>er competit<strong>or</strong><br />

of resources compared to other branches.<br />

Damage to Scots Pine Sinks and Sources<br />

In Scots pine, damage to apical buds has similar effects as in birch (Table 1), and we assume that these effects also<br />

can be explained in the same way (Honkanen et al. 1994). However, the effects of defoliation are m<strong>or</strong>e complex than in the<br />

case of the mountain birch. This can be explained by the simultaneous existence of several needle generations which are in<br />

different positions on a shoot, and which partially feed different meristems (Fig. 1). Defoliation of apical needles in the oneyear-old<br />

needle-class that supply the developing current year foliage early in the season, decreases the growth of new shoots.<br />

However, defoliation of older needles at the same time does not affect the growth of new shoots, presumably because older<br />

needles preferentially supply lateral meristems of the tree (Ericsson 1978, Honkanen and Haukioja MS). In contrast to birch, 5


late season defoliation decreases the growth potential of next year's pine needles m<strong>or</strong>e than early season defoliation<br />

(Honkanen et al. 1994). This presumably relates to the long developmental time of pine buds (Agren and Axelsson 1985)<br />

and that defoliation removes the local st<strong>or</strong>age in pine needles. We assume that late season defoliation of birch mainly hinders<br />

provisioning of the general resource pool (Fig. 1).<br />

In conclusion, both the type of the damaged tissue andtiming of the damage contribute to the post-defoliation<br />

responses of mountain birch and Scots pine. The imp<strong>or</strong>tant points are whether the damaged tissue was a physiological sink<br />

<strong>or</strong> a source, and which sinks the damaged foliage was provisioning at the time of the damage. Thus IA, DIR, RIR, as well as<br />

local changes in growth activity, can all be understood as outcomes of the same basic explanation.<br />

DISCUSSION<br />

The system of sink-source gradients is part of the mechanism regulating the allocation of resources within a plant.<br />

Theref<strong>or</strong>e, it offers a basis f<strong>or</strong> explanations about spatial and temp<strong>or</strong>al patterns of resources, and theref<strong>or</strong>e also f<strong>or</strong> chemical<br />

compounds which are potentially imp<strong>or</strong>tant f<strong>or</strong> herbiv<strong>or</strong>es and about changes in these compounds after herbiv<strong>or</strong>y. This is not<br />

to say that we regard resource availability unimp<strong>or</strong>tant; the availability of carbon and mineral nutrients obviously represents<br />

modifying and constraining fact<strong>or</strong>s f<strong>or</strong> tree-herbiv<strong>or</strong>e interactions. Among other things, resources modify sink strengths of<br />

meristems and plant apical dominance (Cline 1991).<br />

The G/D and, especially, the C/N hypotheses, emphasize plant quality and different ways that sh<strong>or</strong>tages of resources<br />

may cause chemical changes which are assumed to be imp<strong>or</strong>tant f<strong>or</strong> herbiv<strong>or</strong>es. The S/S hypothesis, instead, emphasizes the<br />

ability of meristems to manufacture new biomass via spatially and temp<strong>or</strong>ally variable allocation of resources to specific<br />

meristems. Limitations in resource allocation may alter foliage quality either via the primary <strong>or</strong> the secondary leaf chemistry,<br />

<strong>or</strong>, most probably, both. The SIS hypothesis does not make a pri<strong>or</strong>i predictions of the imp<strong>or</strong>tance of these chemical changes<br />

f<strong>or</strong> herbiv<strong>or</strong>es.<br />

We assume that the basic explanation f<strong>or</strong> DIR, and perhaps f<strong>or</strong> localized RIR, is the weakened state of meristems.<br />

This leads to a lower competitive ability in the plant sink-source system. Altered resistance is an automatic outcome of these<br />

changes. There are four aspects of induced tree responses which supp<strong>or</strong>t <strong>this</strong>. First, defoliation/clipping-induced changes in<br />

plant growth activity are basically local phenomena. This also refers to the largely unspecific changes in primary and<br />

secondary chemistry which produce effects classified as RIR, DIR and IA in trees like birches. This clearly contrasts with<br />

some damage-induced systemic responses which effectively spread within and among plants (Walker-Simmons and Ryan<br />

1977, Farmer and Ryan 1990, Takabayashi et al. 1991). They concern well known and obviously strictly defensive traits like<br />

production and transfer of the proteinase inhibit<strong>or</strong> inducing fact<strong>or</strong> (PIIF) e.g. in many herbaceous plants (McFarland and<br />

Ryaa 1974), <strong>or</strong> the resin duct system of conifers (e.g., Blanche et al. 1992). Second, the responses of birch foliage quality to<br />

defoliations <strong>or</strong> clippings is of a general nature and affects all species of chewing insect herbiv<strong>or</strong>es tested so far in a similar<br />

way (Hanhimfiki 1989). Third, the existence of IA demonstrates that reasons other than defense (<strong>or</strong> recovery) must be sought<br />

when interpreting plant responses to herbiv<strong>or</strong>y. Such seemingly nonadaptive responses are real and presumably reflect<br />

constraints in plant design. Fourth, birch and pine are surprisingly insensitive to the amount of lost biomass: limited as well<br />

as intensive defoliations cause fairly similar degrees of responses (Haukioja and Neuvonen 1987, Honkanen and Haukioja<br />

1994). This is consistent with sink strength regulated photosynthesis which is common in plants (e.g., Wardlaw 1990). All<br />

these aspects indicate that defoliation-induced changes in carbon-based potentially defensive chemistry, as well as in the<br />

primary chemistry, have explanations other than defense, <strong>or</strong> the relative availability of different types of resources. Note that<br />

acceptance of <strong>this</strong> explanation does not deny the imp<strong>or</strong>tance of plant chemistry on herbiv<strong>or</strong>es. However, the question of<br />

whether induced tree resistance, measured as a decrease in insect perf<strong>or</strong>mance after previous foliar damage, is a true defense<br />

<strong>or</strong> not, is outside the scope of <strong>this</strong> paper.<br />

The S/S hypothesis has some obvious practical implications. F<strong>or</strong> instance, branch <strong>or</strong> ramet specific responses of trees<br />

to defoliations (Haukioja et al. 1983, Tuomi et al. 1988b, L_ngstr6m et al. 1990, Hanhim_iki and Senn 1992, Honkanen et al.<br />

1994, Honkanen and Haukioja 1994) become easily comprehensible. This fact has made it possible to use individual ramets<br />

of the same multi-stemmed tree as units on which different treatments have been applied. The question of whether such<br />

branch-specific defoliations are as effective in inducing induced resistance as m<strong>or</strong>e extensive tree-wide defoliations has an<br />

ahnost paradoxical answer: at least in pine, branches show stronger, not weaker, responses after branch-wide than after treewide<br />

defoliations (Honkanen and Haukioja 1994). This result agrees with the S/S hypothesis and, to our understanding,<br />

6


cannot be derived from any other hypothesis: within the canopy, competition of resources among meristems puts a single<br />

branch with defoliation-weakened meristems into a po<strong>or</strong>er relative position if all other branches are intact than if they also are<br />

i weakened. Second, the S/S hypothesis provides a simple explanation f<strong>or</strong> the multiannual "mem<strong>or</strong>y" of DIR (Haukioja el al.<br />

i 988): it might simply result from the time :needed to rest<strong>or</strong>e the vig<strong>or</strong> of meristems. Furtherm<strong>or</strong>e, the implications of the S/<br />

S hypothesis also make it w<strong>or</strong>thwhile to study whether plants have evolved mechanisms to prevent herbiv<strong>or</strong>es causing IA.<br />

F<strong>or</strong> instance, it has been rep<strong>or</strong>ted that developing current year foliage of Scots pine produces 13-keto-8(14)-podocarpen-18oic<br />

acid which deters pine sawflies and decreases use of the young current year foliage (Niemel_i et al. 1982). A strong<br />

deterrence of herbiv<strong>or</strong>es from such current year foliage may reduce the danger of induced ameli<strong>or</strong>ation.<br />

Perhaps the most imp<strong>or</strong>tant empirical message of the S/S hypothesis is a call f<strong>or</strong> extreme caution in damage simulations.<br />

F<strong>or</strong> instance, when triggering the DIR in the foliage of the mountain birch, even small variations in timing (perhaps a<br />

day <strong>or</strong> two) of defoliations may shift the treatment from sink leaves to source leaves. That may alter the response from IA to<br />

RIR <strong>or</strong> DIR, i.e., even the direction of the response. We doubt whether any study published so far has rep<strong>or</strong>ted both the<br />

timing of defoliations in relation to tree phenology, and the exact defoliation practice, in such detail that it would allow the<br />

treatments to be precisely repeated. Actually, we are lacking a proper classification of damages to plants. Theref<strong>or</strong>e, it is not<br />

astonishing that rep<strong>or</strong>ts of plant responses to defoliations are not<strong>or</strong>iously variable (Lawton 1991). Although insects and<br />

plants may differ with reference to species, populations and even individuals, we regard the main problem the superficial<br />

knowledge of the system under study: we have to understand and to be able to rec<strong>or</strong>d what we are doing with our own trees<br />

when we simulate herbiv<strong>or</strong>e damage.<br />

SUMMARY<br />

The debate about damage-induced changes in food plant quality has concentrated on their active vs incidental<br />

defensive roles. We show f<strong>or</strong> birch and pine that the feedback consequences f<strong>or</strong> herbiv<strong>or</strong>es vary from fav<strong>or</strong>able to harmful<br />

depending on the treated plant part and on timing of the damage. Our results cannot be directly derived from the nutrient<br />

status of the tree. Instead, damage-induced alterations in meristematic sink strength provides a general explanation. In birch,<br />

damage to dominant sinks allows other parts to obtain a greater than n<strong>or</strong>mal share of resources which leads to their vig<strong>or</strong>ous<br />

growth and better tbrage quality f<strong>or</strong> herbiv<strong>or</strong>es. However, damage to young source leaves that are supplying local meristems<br />

f<strong>or</strong> next year, weakens such meristems and reduces their success in competing f<strong>or</strong> resources. This leads to less vig<strong>or</strong>ous<br />

growth and po<strong>or</strong>er plant quality f<strong>or</strong> herbiv<strong>or</strong>es. Damage to mature leaves presumably disturbs provisioning of the general<br />

resource pool of the tree but such effects, especially if they are monit<strong>or</strong>ed locally, may be hard to reveal. An emphasis on<br />

damage-induced effects on plant sinks and sources makes induced resistance, induced ameli<strong>or</strong>ation, as well as localization of<br />

responses m<strong>or</strong>e understandable.<br />

ACKNOWLEDGEMENTS<br />

We thank all the persons who have helped in the practical w<strong>or</strong>k on which <strong>this</strong> paper is based, as well as those who<br />

have helped in f<strong>or</strong>mulating the ideas and who have criticized earlier versions of the present paper (Sinikka Hanhim_iki, Seppo<br />

Neuvonen, Pekka Niemelfi, Matti Rousi, Kai Ruohom_.ki, Janne Suomela, John Vranjic, and Timo Vu<strong>or</strong>isalo). John also<br />

checked our English. Financially <strong>this</strong> study has been made possible by grants from the Academy of Finland, Maj and Tot<br />

Nessling Foundation, and the Lapland F<strong>or</strong>est Damage Project.<br />

LITERATURE CITED<br />

*GREN, G.I. and AXELSSON, B. 1985. C<strong>or</strong>relations between needle and shoot growth characteristics in Scots pine (Pinus<br />

sylvestris L.) stands. F<strong>or</strong>estry 58: 199-205.<br />

BENZ, G. 1974. Negative Riickkoppelung durch Raum- und Nahrungskonkurrenz sowie zyklische VerSnderung dr<br />

Nahrungsgrundlage als Regelprinzip in der Populationsdynamik des Grauen L_irchenwicklers, Zeiraphera diniana<br />

(Guende) (Lep., T<strong>or</strong>tricidae). Z. ang. Ent. 76: 196-228.<br />

BJC)RKMAN, C., LARSSON, S., and GREE R. 1991. Effects of nitrogen fertilization on pine needle chemistry and sawfly<br />

perf<strong>or</strong>mance. Oecologia 86: 202-209. 7


BLANCHE, C.A., LORIO, EL., Jr., SOMMERS, R.A., HODGES, J.D., and NEBEKER, T.E. 1992. Seasonal cambial<br />

growth and development of loblolly pine: xylem f<strong>or</strong>mation, inner bark chemistry, resin ducts and resin flow. F<strong>or</strong>est.<br />

Ecol. Manage; 49:151-165.<br />

BRYANT, J.E, CHAPIN, F.S., III, and KLEIN, D.R. 1983. Carbon/nutrient balance of b<strong>or</strong>eal plants in relation to vertebrate<br />

herbiv<strong>or</strong>y. Oikos 40: 357-368.<br />

BRYANT, J.E, KUROPAT, EJ., REICHARDT, EB., and CLAUSEN, T.E 1991. Controls over the allocation of resources by<br />

woody plants to chemical antiherbiv<strong>or</strong>e defense, p. 83-102. In Palo, R.T. and Robbins, C.T., yds. Plant Defenses<br />

Against Mammalian Herbiv<strong>or</strong>es. CRC Press, Boca Raton, Fl<strong>or</strong>ida.<br />

BRYANT, J.E, REICHARDT, EB., CLAUSEN, T.E, and WERNER, R.A. 1993. Effects of mineral nutrition on delayed<br />

inducible resistance in Alaska paper birch. Ecology 74: 2072-2084.<br />

CLINE, M.G. 1991, Apical dominance. Bot. Rev. 57:318-358.<br />

DANELL, K. and HUSS-DANELL, K. 1985. Feeding by insects and hares on birches earlier affected by moose browsing.<br />

Oikos 44: 75_81.<br />

EDENIUS, L. 1993. Browsing by moose on Scots pine in relation to plant resource availability. Ecology 74:2261-2269.<br />

ERICSSON, A. 1978i Seasonal changes in translocation of 14Cfrom different age-classes of needles of 20-year-old Scots<br />

pine trees (Pinus sylvestris). Physiol. Plant. 43: 351-358.<br />

FARMER, E.E. and RYAN, C.A. 1990. Interplant communication: Airb<strong>or</strong>ne methyl jasmonate induces synthesis of proteinuse<br />

inhibit<strong>or</strong>s in plant leaves. Proc. Natl. Acad. Sci USA 87: 7713-7716.<br />

HANHIM,g, KI, S. I989i Induced resistance in mountain birch: defence against leaf-chewing insect guild and herbiv<strong>or</strong>e<br />

competition: Oecologia 81: 242-248.<br />

HANHIMJkKI, S. and SENN, J. 1992. Sources of variation in rapidly inducible responses to leaf damage in the mountain<br />

birch-insect herbiv<strong>or</strong>e system. Oecologia 91:318-331.<br />

HARTLEY, S.E. and LAWTON, J.H. 1991. Biochemical aspects and significance of the rapidly induced accumulation of<br />

phenolics in birch foliage, p. 105-132. In Tallamy, D.W. and Raupp, M.J., yds. Phytochemical Induction By Herbiv<strong>or</strong>es.<br />

John Wiley, New Y<strong>or</strong>k.<br />

HAUKIOJA, E. 198Z Inducible defences of white birch to a geometrid defoliat<strong>or</strong>, Epirrita autumnata, p. 199-203. In<br />

Visser, J.H, and Minks, A.K., yds. Insect-Plant Relationships. PUDOC, Wageningen.<br />

HAUKIOJA, E. I991a! Cyclic fluctuations in density: interactions between a defoliat<strong>or</strong> and its host tree. Acta Oecol. 12:<br />

7-88<br />

HAUKIOJA, E. 1991b_ The influence of grazing on the evolution, m<strong>or</strong>phology and physiology of plants as modular<br />

<strong>or</strong>ganisms. Philos. Trans. R. Soc. Loud. B. 333: 241-247.<br />

HAUKIOJA, E. and NEUVONEN , S. 1985. Induced long-term resistance in birch foliage against defoliat<strong>or</strong>s: defensive <strong>or</strong><br />

incidental? E_61ogy 66:1303-1308.<br />

HAUKIOJA, E. and NEUVONEN, S. 1987. Insect population dynamics and induction of plant resistance: the testing of<br />

hypotheses, p _11'432. In Barbosa, E and Schultz, J.C., yds. Insect Outbreaks. Academic Press, San Diego.<br />

HAUKIOJA, E., KAPIAINEN, K., NIEMELA, P., and TUOMI, J. 1983. Plant availability hypothesis and other explana-<br />

8


HAUKIOJA, E., SUOMELA, J., and NEUVONEN, S. t985. Long-term inducible resistance in birch foliage: triggering<br />

cues and efficacy on a defoliat<strong>or</strong>. Oecologia 65: 363-369.<br />

HAUKIOJA, E., NEUVONEN, S., HANHIMAKI, S., and NIEMEL._, R 1988. The autumnal moth in Fennoscandia, p.<br />

163-178. In Berryman, A.A., ed. Dynamics of F<strong>or</strong>est Insect Populations. Patterns, Causes, Implications. Plenum<br />

Press, New Y<strong>or</strong>k.<br />

HAUKIOJA, E., RUOHOMAKI, K., SENN, J., SUOMELA, J., and WALLS, M. 1990. Consequences of herbiv<strong>or</strong>y in the<br />

mountain birch (Betula pubescens ssp t<strong>or</strong>tuosa): imp<strong>or</strong>tance of the functional <strong>or</strong>ganization of the tree. Oecologia 82:<br />

238- 247.<br />

HERMS, D.A. and MATTSON, W.J. 1992. The dilemma of plants: to grow <strong>or</strong> defend. Quart. Rev. Biol. 67: 283-335.<br />

HONKANEN, T. and HAUKIOJA, E. 1994. Why does a branch suffer m<strong>or</strong>e after branch-wide than after tree-wide defoliation?<br />

Oikos 71 : 441-450.<br />

HONKANEN, T., HAUKIOJA, E., and SUOMELA, J. 1994. Effects of simulated defoliation and debudding on needle and<br />

shoot growth in Scots pine (Pinus syIvestris): implications of plant source/sink relationships f<strong>or</strong> plant-herbiv<strong>or</strong>e<br />

studies. Funct. Ecol. 8: 631-639.<br />

IASON, G. and HESTER, A.J. 1993. The response of heather (Calluna vulgaris) to shade and nutrients - predictions of the<br />

carbon-nutrient balance hypothesis. J. Ecol. 81: 75-80.<br />

KOZLOWSKI, T.T. 1971. Growth and Development of Trees. Academic Press, New Y<strong>or</strong>k.<br />

LA.NGSTROM, B., TENOW, O., ERICSSON, A., HELLQVIST, C., and LARSSON, S. 1990. Effects of shoot pruning on<br />

stem growth, needle biomass, and dynamics of carbohydrates and nitrogen in Scots pine as related to season and tree<br />

age. Can. J. F<strong>or</strong>. Res. 20: 514-523.<br />

LAWTON, J. 1991. Ecology as she is done, and could be done. Oikos 61: 289-290.<br />

LOYTTYNIEMI, K. 1985. On repeated browsing of Scots pine saplings by moose (Alces alces). Silva Fenn. 19: 387-391.<br />

MCFARLAND, D. and RYAN, C.A. 1974. Proteinase inhibit<strong>or</strong>-inducing fact<strong>or</strong> in plant leaves: a phylogenetic survey. Plant<br />

Physiol. 54: 706-708.<br />

NEUVONEN, S. and DANELL, K. 1987. Does browsing modify the quality of birch foliage f<strong>or</strong> Epirrita autumnata larvae?<br />

Oikos 49: 156-160.<br />

NEUVONEN, S. and HAUKIOJA, E. 1991. The effects of inducible resistance in host foliage on birch-feeding herbiv<strong>or</strong>es,<br />

p. 277-291. In Tallamy, D.W. and Raupp, M.J., eds. Phytochemical Induction By Herbiv<strong>or</strong>es. John Wiley and Sons,<br />

New Y<strong>or</strong>k.<br />

NEUVONEN, S., HANHIMA.KI, S., SUOMELA, J., and HAUKIOJA, E. 1988. Early season damage to birch foliage<br />

affects the perf<strong>or</strong>mance of a late season herbiv<strong>or</strong>e. J. Appl. Entomol. 105: 182-189.<br />

NIEMEL,_., R, ARO, E.-M., and HAUKIOJA, E. 1979. Birch leaves as a resource f<strong>or</strong> herbiv<strong>or</strong>es. Damage-induced increase<br />

in leaf phenols with trypsin-inhibiting effects. Rep. Kevo Subarct. Res. Stat. 15: 37-40.<br />

NIEMEL*, R, MANNILA, R., and M,_.NTSA.LA, R 1982. Deterrent in Scots pine, Pinus sylvestris, influencing feeding<br />

behavi<strong>or</strong> of the larvae of Neodiprion sertifer (Hymenoptera, Diprionidae). Ann. Ent. Fenn. 48: 57-59.<br />

NIEMELJ_, R, TUOMI, J., MANNILA, R., and OJALA, R 1984. The effect of previous damage on the quality of Scots<br />

pine foliage as food f<strong>or</strong> Diprionid sawflies. Z. Ang. Ent. 98: 33-43.


NIEMELA, R, TUOMI, J,, and LOJANDER, T. 1991. Defoliation of the Scots pine and perf<strong>or</strong>mance of diprionid sawflies.<br />

J. Anim. Ecol! 60! 683-692.<br />

NUORTEVA, H. and KURKELA, T. 1993. Effects of crown reduction on needle nutrient status of scleroderris-cankerdiseased<br />

and greefi,pruned Scots pine. Can. J. F<strong>or</strong>. Res. 23:1169- 1178.<br />

REICH, RB., WALTERS; M.B., KRAUSE, S.C., VANDERKLEIN, D.W., RAFFA, K.F., and TABONE, T. 1993. Growth,<br />

nutrition and gas exchange of einus resinosa followingartificial defoliation. Trees 7: 67-77.<br />

ROSSITER, M.C., SCHULTZ, J.C., and BALDWIN, I.T. 1988. Relationships among defoliation, red oak phenolics, and<br />

gypsy moth growth and reproduction. Ecology 69: 267-277.<br />

RUOHOMAKI, K.i eHAPIN, F. S., HAUKIOJA, E., NEUVONEN, S. and SUOMELA, J. Delayed inducible resistance in<br />

mountain birch in resp°nse to fertilization and shade. Ecology. In press.<br />

SACHS, T., NOVOPLANSKY, A., and COHEN, D. 1993. Plants as competing populations of redundant <strong>or</strong>gans. Plant Cell<br />

Environ. 16:765-770.<br />

SENN, J. and 1994. Reactions of the mountain birch to bud removal: effects of severity and timing, and<br />

implications f<strong>or</strong> herbiv<strong>or</strong>es. Funct. Ecol. 8: 494-501.<br />

SPRUGEL, D.G.,HINCKLFX, %M.,<br />

Ecol. Syst. 221309;334.<br />

,,d SCHAAP, W. 1991 The the<strong>or</strong>y and practice of branch autonomy, annu. Rev.<br />

SUTHERLAND, W. and STILLMAN, R.A. 1990. Clonal growth: insights from models, p. 95-111. In van Groenendael, J.<br />

and de Krooni H_i eds. Clonal Growth in Plants: Regulation and Function. SPB Academic Publishing, The Hague.<br />

10<br />

vian mountaifi eh_h and n<strong>or</strong>thern Finland 1862-1968. Zool. Bidr. Uppsala 2: 1-107.


DAMAGE-INDUCED NUTRITIONAL CHANGES IN PINE FOLIAGE:<br />

AN OVERVIEW<br />

PAIVI LYYTIK_INEN<br />

Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Department of F<strong>or</strong>est Ecology<br />

RO. Box 18, FIN-01301 Vantaa, Finland<br />

INTRODUCTION<br />

Plant resistance can be divided into constitutive and induced defenses. Constitutive defense can deter, repel, intoxi-<br />

cate <strong>or</strong> interfere with insect herbiv<strong>or</strong>es° This can change in response to external <strong>or</strong> internal influences (e.g., weather, site<br />

fact<strong>or</strong>s, plant age), bt_t is t.lsua]ly stable (Schtdt 1988). Induced responses are a f<strong>or</strong>m of phenotypic plasticity and can shift<br />

plant resistance as a result of herbiv<strong>or</strong>e attack (Haukioja and Neuvonen 1987, Haukioja 1990). Induced changes in foliage<br />

quality can take place in a density-dependent fashion and modify the stability of insect populations. Responses which are<br />

caused by one insect generation and act during subsequent generations have been called delayed induced resistance (DIR).<br />

Those responses which _)ccur in the same insect generation have been called rapid induced resistance (RIR) (Haukioia and<br />

Neuvonen 1987_ Hauki_4ja 1990). Emphasis has usually been placed on the defensive role of chemical compounds, leading to<br />

underestimation of the ro_e of some *_positive" compounds (e.g,, mineral nutrients, water, sugars). It is often questioned<br />

whether chemical changes (low quality to herbiv<strong>or</strong>es) represent an active defense <strong>or</strong> are only incidental by-products <strong>or</strong> reflect<br />

the: abiotic environment (Tuomi et al. 1984, Haukioja and Neuvonen 1985).<br />

Plant species <strong>or</strong> individuals may respond differently to insect-caused <strong>or</strong> artificial def_._liation. The induced responses<br />

of deciduous trees have been studied m<strong>or</strong>e than those of evergreen trees. The most studied tree genera are Bemla, Quercus,<br />

A[nus and SMix (Schu]t and Baldwin 1982, Jeker 1983, Raupp and Denno 1984, Haukioja and Neuvonen 1987, Neuvonen et<br />

al. 1987). Defoliation by Zeiraphera diniana altered the contents of nitrogen and fiber in European larch, Lc_rix decidua, f<strong>or</strong><br />

a four-year period (Baltensweiler et al. 1977). Responses to insect defoliation have been detected in the foliage of Pinus<br />

ponderosa (Wagner and Evans 1985), P cont<strong>or</strong>ta (Leather et al. 1987) and P sylvestris (Lyytik_iinen 1993). Induced reac-<br />

tions have also been rep<strong>or</strong>ted in P. radiata after damage by spider mites (Karban 1990).<br />

A very imp<strong>or</strong>tant question in insect ecology is whether certain tree species can exhibit induced responses and whether<br />

such responses have an impact on defoliating insect pests. The main aim of <strong>this</strong> study is to find those circumstances when<br />

induced responses have been expressed and affected insect perf<strong>or</strong>mance <strong>or</strong> perhaps population dynamics. The discussion is<br />

focused on induced reactions in pine foliage and experiments with defoliating insects.<br />

QUALITATIVE CHANGES IN PINES<br />

Rather few studies have been carried out on real <strong>or</strong> artificial defoliation in conifers, especially pines. The very first<br />

investigations dealt with modified polyphenol metabolism in P. sylvestris after attack by Neodiprion sertifer (Thielges 1968),<br />

but the conclusions were erroneous due to indirect comparison between the occurrence of the compound and sawfly attack.<br />

Some of the rep<strong>or</strong>ted studies have involved insect rearings without chemical analyses, only leading to hypotheses about the<br />

causal fact<strong>or</strong>s. One difficulty is the wide variability in experimental design, which makes comparisons difficult. One basic<br />

problem seems to be the analysis of needle year-classes of different age. The nutritive status, i.e., the value of foliage as<br />

photosynthetic machinery and the ability to react after damage, are not the same in different year-classes (Bernard-Dagan<br />

1988, Haukioja 1990). The timing of treatments during the growing season can also affect the expression of the reactions<br />

(Hartley and Lawton 1991).<br />

The nutrients most studied are nitrogen, minerals, sugars, and water content (Table 1). However, needle resin acids<br />

that are obviously harmful f<strong>or</strong> diprionid sawflies (Larsson et al. 1986) have been analyzed only once after defoliation. Rapid<br />

Mattson, W.J., Niemel_i, E, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest fbr pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

11


induced reactions have been studied in P ponderosa and P. sylves_ris. In P. po_derosa the nitrogen and tannin concentrations<br />

decreased, but those of proteins, phenols, procyanidins and certain nutrients increased in P. sylvestris. Needle water content<br />

tended to rise sh<strong>or</strong>tly after damage. After one <strong>or</strong> m<strong>or</strong>e years (Table 1) the nitrogen, minerals, and monoterpene concentrations<br />

in P. cont<strong>or</strong>ta increased, but only that of P decreased. A similar trend f<strong>or</strong> the nutrient and water contents in P. sylvestris<br />

has also been rep<strong>or</strong>ted. The most interesting aspect is the increase in diterpene acids and sugars, but there are only two<br />

rep<strong>or</strong>ts of such findings (Buratti el al. 1988, Lyytik_iinen 1994).<br />

Artificial and natural defoliation induce either rapid <strong>or</strong> delayed chemical changes in pine foliage. The response to<br />

insect-caused defoliation is usually stronger than that to artificial damage (Neuvonen et al. 1987, Harttey and Lawton 1991),<br />

but opposing reactions have also been rep<strong>or</strong>ted (Baldwin 1988). It would appear that pines are m<strong>or</strong>e insensitive to defoliation<br />

than deciduous trees. The damage level and its timing are probably m<strong>or</strong>e imp<strong>or</strong>tant fact<strong>or</strong>s f<strong>or</strong> the expression of the reactions<br />

than the damage mode. Deciduous trees can respond to low foliage damage (e.g., tearing) (Haukioja and Neuvonen 1987),<br />

but conifers need at least moderate damage (Wagner and Evans 1985, Leather et al. 1987, Lyytikfiinen 1992a). In some cases<br />

the qualitative changes in late summer were different than those iraearly summer (e.g., Lyytik_iinen 1994). Ericson et al.<br />

(t980a,b) observed that Scots pines defoliated late in the season suffered from lowered starch concentrations, and the trees<br />

with the highest degree of damage lost their current year's shoots and new buds. The studies referred to in Table 1 have<br />

mostly been concerned with responses to sawfly dalnage, induced reactions have unf<strong>or</strong>tunately been investigated with very<br />

few other pine insect species, even though there are many other harmful pests on pines in Europe, e.g., Panolisflammea<br />

(Barbour 1987).<br />

NEEDLE QUALITY AND THE PERFORMANCE OF DEFOLIATING INSECTS<br />

Rapid Responses<br />

Qualitative changes do occur in pines, but are these changes targeted at herbiv<strong>or</strong>es and are the species used sensitive<br />

enough to respond to induced reactions? Rapid reactions and insect response have been detected in two pine species: Pinus<br />

ponderosa and P sylvestris (Table 2). Artificial damage adversely affected the growth of Neodiprionfulviceps, which should<br />

indicate negative reactions in P. ponderosa. The response of Neodiprion sertifer was mainly positive on P. sylvestris sh<strong>or</strong>tly<br />

after insect-caused defoliation, but after artificial damage there were no differences between control and experimental trees.<br />

Another sawfly species responded either negatively <strong>or</strong> not at all. Furtherm<strong>or</strong>e, the results with Scots pine were not parallel.<br />

It would appear that Diprion pint was the least sensitive to qualitative changes in the needles. On the other hand, the low<br />

defoliation level f<strong>or</strong> the whole canopy may have an effect on the indifferent perf<strong>or</strong>mance of different sawfly species (Niemel_i<br />

et at. 1984). In general, slight defoliation caused no changes in insect perf<strong>or</strong>mance. In some cases slight defoliation may<br />

even have adverse effects on insect success compared to heavy defoliation (LyytikE.inen 1992a). There were some positive<br />

responses in early summer and soon after damage, but mainly negative responses in late summer. The lack of impact on<br />

sawfly survival indicates no effect on population dynamics. On the basis of these rep<strong>or</strong>ts (Table 2), it would appear that<br />

Scots pine possesses no effective, rapid resistance reactions against diprionid sawflies.<br />

Delayed Responses<br />

M<strong>or</strong>e studies have been carried out on delayed reactions in pines than on rapid reactions. In P. cont<strong>or</strong>ta the investiga-<br />

tions deal only with experiments on seedlings (Table 3). The negative effect of treatment was obvious after defoliation by<br />

Panolisflammea, reflecting the persistence of induced reactions harmful to insects. One possible explanation f<strong>or</strong> <strong>this</strong> may be<br />

that current-year needles were used in experiments and they obviously contain m<strong>or</strong>e secondary compounds than older needle<br />

year-classes (Buratti et al. 1988, 1990). There was a trend f<strong>or</strong> positive <strong>or</strong> neutral responses after previous early summer<br />

defoliation. However, the responses were mostly negative following late summer damage in both pine species (Table 3). The<br />

results f<strong>or</strong> P flammea only are dissimilar due to the different experimental design. In general the results f<strong>or</strong> P sylvestris were<br />

negative <strong>or</strong> neutral. There were also experiments with N. sertifer and Gilpinia pallida in which there were no responses. As<br />

in studies on rapid responses, delayed responses did not usually affect larval survival, thus indicatingnegligible effects on<br />

population dynamics.<br />

The imp<strong>or</strong>tance of defoliation level was also obvious in the rep<strong>or</strong>ts of delayed induced reactions in pines. The<br />

direction of the response was not similar at different levels of defoliation (Lyytik_iinen 1993, 1994), <strong>or</strong> else the response in<br />

sawfly perf<strong>or</strong>mance only occurred after moderate damage (Lyytik_iinen 1992a). In some cases, the sawfly response was<br />

13


nonlinear (Lyytik_iinen 1992a), indicating some kind of optimality perhaps in nutrient status due to stress (Wagner and Frant<br />

1990). The results in these rep<strong>or</strong>ts (Table 3) testify that lodgepole pine obviously responds negatively to defoliating insects,<br />

but in Scots pine the responses are weaker and are only expressed under certain conditions.<br />

DISCUSSION<br />

Foliage of different ages may respond differently to defoliation. Several rep<strong>or</strong>ts indicate that qualitative changes<br />

induced after spring defoliation in deciduous trees are m<strong>or</strong>e obvious those occurring after late summer defoliation (e.g.,<br />

Haukioja and Niemelfi 1979). Young, expanding foliage acts as a sink f<strong>or</strong> metabolites, and plant resources are converted to<br />

foliage development and expansion (Coleman 1986). Maj<strong>or</strong> changes in the biochemistry and physiology can occur as a result<br />

of foliage damage. On the other hand, damage to older foliage may decrease the amount of resources available to young<br />

foliage (Coleman 1986). However, no studies have been carried out on the gradual changes in needle quality after detbliation<br />

taking place during successive growing season(s). Needle quality is usually based on a single sampling. In Pinus pinaster,<br />

sugar, lipid, starch, and terpene concentrations have been found to vary acc<strong>or</strong>ding to season (Bernard-Dagan 1988). This<br />

phenomenon may offer one explanation f<strong>or</strong> the findings concerning different responses following early <strong>or</strong> late summer<br />

defoliation.<br />

Plant resource allocation and partitioning change in relation to age. Seedlings allocate most of their resources to<br />

growth, and mature plants allocate to reproductive <strong>or</strong>gans (Kolowski 1971). Induced reactions have been found in three pine<br />

species: P. cont<strong>or</strong>ta, P ponderosa and P radiata (Leather et al. 1987, Wagner 1988, Karban 1990). The growth of P. radiata<br />

saplings is the fastest (Kolowski 1971). P. ponderosa and P. cont<strong>or</strong>ta are regarded as slow-growing species, but the growth<br />

of P. sylvestris iseven slower than that of R cont<strong>or</strong>ta (Lyytik_iinen 1993). The lack of induced reactions is common in<br />

inherentlyslow-growing conifers, relying on constitutive resistance (Rhoades and Cates 1976). The induced reactions<br />

detected inthese pinespecies were linked with age and fast growth because seedlings and saplings primarily showed the<br />

responses. In lodgepole pine, trees over 20 years old showed no induced resistance against herbiv<strong>or</strong>es (Leather et al. 1987,<br />

Watt et al. 1991), and only at ages > 12 years, did saplings suffer from outbreaks of Panolis flammea (Barbour 1987). In<br />

Scots pine the reactions were m<strong>or</strong>e pronounced in seedlings than in saplings (Lyytik_,inen 1992b, 1993, 1994). The defensive<br />

strategy of Scots pine is obviously also age-specific, i.e., the seedlings resemble fast-growing species in showing induced<br />

responses.<br />

Plants usually show flexibility in their physiology within the bounds set by environmental limitations (e.g., drought,<br />

shading, nutrient deficiency). Changes in resource allocation may be directed towards equalization of all the fact<strong>or</strong>s limiting<br />

plant growth (e.g., Hunt and Nicholls 1986). Environmental stresses may cause shifts in the partitioning of resources to<br />

et al. 1987). Bryant et al. (1983) suggested that the phenotypic responses of secondary chemistry<br />

are carbon-nutrient balance. Carbon-based, secondary metabolites are supposed to accumulate in<br />

is limited by mineral nutrients (Tuomi et al. 1988, Tuomi 1992). The defense bears no cost, because<br />

supp<strong>or</strong>ted by the resources acquired in excess of primary metabolic requirements (Tuomi et al.<br />

lbundant nutrient status of the growing site in many studies could have reduced the probability of<br />

c production of secondary metabolites (e.g., Niemel_i et al. 1991, Lyytik_iinen 1993, 1994).<br />

f<strong>or</strong> only weak rapid <strong>or</strong> delayed reactions in pines may be due to physiological differences between<br />

trees. In deciduous trees the stems and roots provide the maj<strong>or</strong> carbon reserves (Bryant et al. 1983).<br />

In and current photosynthates in evergreen trees are st<strong>or</strong>ed in old foliage (Tuomi et al. 1984, 1988),<br />

e.g., by sawfly larvae. These physiological differences may contribute to the different pattern of<br />

Ion (Niemel_i and Tuomi 1993) because of a sh<strong>or</strong>tage of mineral nutrients <strong>or</strong> carbon (Niemel_i et al.<br />

1984)_<br />

used in the experiments covered here did not respond in a similar manner to damage. Defoliation<br />

of and saplings (Table 3) affected the success of G. pallida, but not that of IV.sertifer (Lyytikfiinen<br />

the effect of damage affected the perf<strong>or</strong>mance of R flammea on P cont<strong>or</strong>ta (Leather et al. 1987) and<br />

N. (Wagner 1986). One explanation f<strong>or</strong> these findings could be the different sensitivity of the<br />

(Hanski 1987, Haukioja 1990, Niemel_i and Tuomi 1993). Outbreak species such as N. sertifer may<br />

be :es in food quality than less common species, e.g., G. pallida on the same host species. Another<br />

ex in larval feeding periods. The larvae of N. sertifer consume only mature needles in early


summer, but the other sawfly species can also eat the current foliage during rniddle <strong>or</strong> late summer. Needle year-classes may<br />

also react differently after needle damage (Niemel_ and Tuomi 1993), <strong>or</strong> interactions among needle year-classes may occur<br />

(Buratti et al. 11988). Thirdly, as earlier rnentioned, there could also be phenological changes in the same year-classes during<br />

the summer (Bernard-Dagan 1988). The quality of mature needles is different m early season species compared to late<br />

season species.<br />

The existence of induced responses in P. cont<strong>or</strong>ta and R potzderosa may in part ofter one explanation f<strong>or</strong> the population<br />

dynamics of the species feeding on them (e.g., Barbour 1988). In Scots pine the weak induced reactions may be the<br />

reason f<strong>or</strong> the tack of cydicity in sawfly outbreaks (Niemela and Tuomi 1993). The lack of damage-induced reactions in the<br />

needles of mature and pole-size Scots pine enables outbreaks of N. sertifer to occur f<strong>or</strong> several years at the same site. The<br />

explanations f<strong>or</strong> the different phases in population dynamics must be due to fact<strong>or</strong>s (e.g., parasitoids, diseases) other than<br />

qualitative changes in the foliage.<br />

SUMMARY<br />

In pine foliage, induced responses against defoliating insects have been detected in P#zus ponderosa, P. cont<strong>or</strong>ta and<br />

P. syh,estris. Contents of nitrogen, minerals, and water tended to increase after damage. The damage level and timing are<br />

imp<strong>or</strong>tant fact<strong>or</strong>s f<strong>or</strong> the expression of these reactions. P. ponderosa showed evidence of rapid induced reactions, but not in<br />

E sylvestris. Delayed induced reactions were obvious in P. cotzt<strong>or</strong>ta, but in P. sylvestris the responses were weaker and<br />

expressed only under certain conditions. Foliage of different ages may respond differently, because young foliage acts early<br />

on as a sink f<strong>or</strong> metabolites. Furtherm<strong>or</strong>e, foliage quality can change acc<strong>or</strong>ding to season. The induced reactions detected in<br />

certain pine species are linked to their seedlings and fast growth. The defensive strategy of pines is probably highly agespecific:<br />

seedlings resemble fast-growing species. The explanation fbr the typically weak reactions in pines may be due to<br />

physiological differences between them and deciduous trees. The weak induced resistance in Scots pine may explain lack of<br />

cyclicity in pine sawfly outbreaks.<br />

ACKNOWLEDGMENTS<br />

I wish to thank Pekka Niemel_i _br his critical and helpful comments on the manuscript and John Derome f<strong>or</strong> kindly<br />

revising the English text. The w<strong>or</strong>k was financially supp<strong>or</strong>ted by The Academy of Finland (project no. 209103 1).<br />

LITERATURE CITED<br />

BALDWIN, I.T. 1988. The alkaloidal response of wild tobacco to real and simulated herbiv<strong>or</strong>y. Oecologia 77: 378-381.<br />

BALTENSWEILER, W., BEN, G., BOVEY, R and DELUCCHI, V. 1977. Dynamics of larch bud moth populations. Annu.<br />

Rev. Entomol. 22: 79-100.<br />

BARBOUR, D.A. 1987. Pine beauty moth population dynamics. General considerations and life-table w<strong>or</strong>k. In Leather,<br />

S.R., Stoakley, J.T., and Evans, H.E, eds. Population biology and control of the pine beauty moth. F<strong>or</strong>. Comm. Bull.<br />

67: 7-13.<br />

BARBOUR, D.A. 1988. The pine looper in Britain and Europe, p. 291-308. In Berryman, A.A., ed. Dynamics of F<strong>or</strong>est<br />

Insect Populations: Patterns, Causes, and Management Strategies. Plenum Press, New Y<strong>or</strong>k.<br />

BERNARD-DAGAN, C. 1988. Seasonal variations in energy sources and biosynthesis of terpenes in maritime pine, p. 93-<br />

115. In Mattson, W.J., Levieux, J., and Bernard-Dagan, C., eds. Mechanisms of Woody Plant Defenses Against<br />

Insects: Search f<strong>or</strong> Pattern. Springer-Verlag, New Y<strong>or</strong>k.<br />

BRYANT, J.R, CHAPIN, F.S.III, and KLEIN, DR. 1983. Carbon/nutrient balance of b<strong>or</strong>eal plants in relation to vertebrate<br />

herbiv<strong>or</strong>y. Oikos 40: 357-368.<br />

17


BURATTI, L., ALLAIS, J.E, and BARBIER, M. 1988. The role of resin acids in the relationships between Scots pine and<br />

the sawfly, Diprion pini (Hymenoptera: Diprionidae). I Resin acids in the needles, p. 171-187. In Mattson, W.J.,<br />

Levieux, J., and Bernard-Dagan, C., eds. Mechanisms of Woody Plant Defenses Against Insects: Search f<strong>or</strong> Pattern.<br />

Springer-Verlag, New Y<strong>or</strong>k.<br />

BURATTI, L., ALLAIS, J.E, GERI, C., and BARBIER, M. 1990. Abietane and pimarine diterpene acid evolution in Scots<br />

pine, Pinus sylvestris, needles in relation to feeding of the pine sawfly, Diprion pini L. Ann. Sci. F<strong>or</strong>. 47:161-171.<br />

CHAPIN, ES.III, BLOOM, A.J., FIELD, C.B,. and WARING, R.H. 1987. Plant responses to multiple environmental<br />

fact<strong>or</strong>s. Bio Science 37: 49-57.<br />

COLEMAN, J.S. 1986. Leaf development and leaf stress: increased susceptibility associated with sink-source transition.<br />

Tree Physiol. 2: 289-299.<br />

ERICSSON, A, HELLKVIST, K., HILLERDAHL-HAGSTROMER, K., LARSSON, S., MATTSON-DJOS, E., and<br />

TENOW, O. 1980a. Consumption and pine growth - hypothesis on effects on growth processes by needle-eating<br />

insects. In Persson, T, ed. Structure and Function of N<strong>or</strong>thern Coniferous F<strong>or</strong>ests - an ecosystem study. Ecol. Bull.<br />

(Stockholm) 32: 537-545.<br />

ERI , A., LARSSON, S., and TENOW, O. 1980b. Effects of early and late season defoliation on growth and carbohydrate<br />

dynamics in Scots pine. J. Appl. Ecol. 17: 747-769.<br />

GERI , L., and ALLAIS, J.E 1988. The role of resin acids in the relationship between Scots pine and sawfly,<br />

Oiprionpini (Hymenoptera: Diprionidae). II. C<strong>or</strong>relations with the biology of Diprion pini, p. 189-201. In Mattson,<br />

and Bernard-Dagan, C., eds. Mechanisms of Woody Plant Defenses against Insects: Search f<strong>or</strong><br />

New Y<strong>or</strong>k.<br />

109: 436_447.<br />

E, and LEVIEUX, J. 1990. Incidence de la consommation du feullage de Pins sylvestres<br />

filles sur le developpement et la fecodite de Diprion pini L. (Hym., Diprionidae). J. Appl. Ent.<br />

HANS,, I. I987. Pine sawfly population dynamics: patterns, processes, problems. Oikos 44: 165-174.<br />

HA_EY, S.E. and LAWTON, J.H. 1991. Biochemical aspects and significance of the rapidly induced accumulation of<br />

foliage, p. 105-132. In Tallamy, D.W. and Raupp, M.J., eds. Phytochemical Induction by Herbi-<br />

',wY<strong>or</strong>k.<br />

HAU Ai E. 1990. Induction of defenses in trees. Annu. Rev. Entomol. 36: 25-42.<br />

HA1 S. 1985. Induced long-term resistance of birch foliage against defoliat<strong>or</strong>s: defensive <strong>or</strong><br />

,gy 66:1303-1308.<br />

18<br />

S. 1987. Insect population dynamics and induction of plant resistance: the testing of<br />

1-432. In Barbosa, E and Schult, J.C., eds. Insect Outbreaks. Academic, San Diego.<br />

A.D. 1986. Stress and coarse control of root-shoot partitioning in herbaceous plants. Oikos 47:<br />

einer Defoliation im V<strong>or</strong>jahr und des Blattalters auf die Larvenentwicklung von Melasoma<br />

)tera, Chrysomelidae). Mitt. Schwei. Ent. Ges. 56: 237-244.<br />

outbreaks on only young trees: testing hypotheses about aging and induced resistance.<br />

Growth and development of trees. Academic Press, New Y<strong>or</strong>k.


LARSSON, S., BJ(_RKMAN, C., and GREF, R. 1986. Responses of Neodiprion sertifer (Hym., Diprionidae) larvae to<br />

variation in needle resin acid concentration in Scots pine. Oecologia 70: 77-84.<br />

LEATHER, S.R., WATT,A.D., and FORREST, G.I. 1987. Insect-induced changes in young lodgepole pine (Pinus cont<strong>or</strong>ta):<br />

the effect of previous defoliation on oviposition, growth and survival of pine beauty moth, Panolisflammea. Ecol.<br />

Entomol. 12: 275-281.<br />

LYYTIKAINEN, R 1992a. The influence of damage level in Pinus sylvestlqs foliage on the perf<strong>or</strong>mance of diprionid<br />

sawflies. Scand. J. F<strong>or</strong>. Res. 7: 249-257.<br />

LYYTIKAINEN, R 1992b. Comparison of the effects of artificial and natural defoliation on the growth of diprionid sawflies<br />

on Scots pine foliage. J. Appl. Ent. 114: 57-66.<br />

LYYTIK.AINEN, R 1993. The perf<strong>or</strong>mance of diprionid sawflies on defoliated pine seedlings. Acta Oecologica 14: 305-<br />

315.<br />

LYYTIKAINEN, P. 1994. Effects of natural and artificial defoliations on sawfly perf<strong>or</strong>mance and foliar chemistry of Scots<br />

pine saplings. Ann. ool. Fennici (In Press).<br />

NEUVONEN, S., HAUKIOJA, E. and MOLARIUS, A. 1987. Delayed inducible resistance against a leaf-chewing insect in<br />

four deciduous tree species. Oecologia 74: 363-369.<br />

NIEMELJ(, P. and TUOMI, J. 1993. Sawflies and inducible resistance of woody plants, p. 391-407. In Wagner, M.R. and<br />

Raffa, K.F. Sawfly Life Hist<strong>or</strong>y Adaptations to Woody Plants. Academic Press, New Y<strong>or</strong>k.<br />

NIEMELJk, E, TUOMI, J., and LOJANDER, T. 1991. Defoliation of the Scots pine and perf<strong>or</strong>mance of diprionid sawflies.<br />

J. Anirn. Ecol. 90: 683-692.<br />

NIEMEL,i_, R, TUOMI, J., MANNILA, R., and OJALA, R 1994. The effect of previous damage on the quality of Scots<br />

pine foliage as food f<strong>or</strong> diprionid sawflies. Angew. Entomol. 98: 33-43.<br />

RAUPP, M.J. and DENNO, R.E 1984. The suitability of damaged willow leaves as food f<strong>or</strong> the leaf beetle, Plagiodera<br />

versicol<strong>or</strong>a. Ecol. Entomol. 9: 443-448.<br />

RHOADES, D.F. and CATES, R.G. 1976. Towards a general the<strong>or</strong>y of plant antiherbiv<strong>or</strong>e chemistry. Rec. Adv. Phytochem.<br />

10: 168-213.<br />

SCHULTZ, J.C. 1988. Plant responses induced by herbiv<strong>or</strong>es. Trends. Ecol. Evol. 3: 45-49.<br />

SCHULTZ, J.C. and BALDWIN, I.T. 1982. Oak leaf quality declines in response to defoliation by gypsy moth larvae.<br />

Science 217: 149-151.<br />

THIELGES, B.A. 1968. Altered polyphenol metabolism in the foliage of Pinus sylvestris associated with European sawfly<br />

attack. Can. J. Bot. 46: 724-725.<br />

TUOMI, J. 1992. Toward integration of plant defence the<strong>or</strong>ies. TREE 7: 365-367.<br />

TUOMI, J., NIEMELA, R, CHAPIN, F.S., BRYANT, J.R, and SIREN, S. 1988. Defensive responses of trees in relation to<br />

their carbon/nutrient balance, p. 57-72. In Mattson, W.J., Levieux, J., and Bernard-Dagan, C., eds. Mechanisms of<br />

Woody Plant Defenses against Insects: Search f<strong>or</strong> Pattern. Springer, New Y<strong>or</strong>k.<br />

TUOMI, J., NIEMEL._, E, HAUKIOJA, E., SIREN, S., and NEUVONEN, S. 1984. Nutrient stress. An explanation f<strong>or</strong><br />

plant antiherbiv<strong>or</strong>e responses to defoliation. Oecologia 61: 208-210.<br />

19


WAGNER, ',nce of moisture stress and induced resistance in ponderosa pine, Pinus ponderosa Dougl. ex<br />

y, Neodiprion autumnaIis Smith. F<strong>or</strong>. Ecol. Manage. 15: 43-53.<br />

83:452-457.<br />

defenses in ponderosa pine against defoliating insects, p. 141-155. In Mattson, W.J.,<br />

gan, C., eds. Mechanisms of Woody Plant Defenses against Insects: Search f<strong>or</strong> Pattern.<br />

RD. 1985. Defoliation increases nutritional quality and allelochemicals of pine seedlings.<br />

P7.<br />

D.R 1990. Influence of induced water stress in ponderosa pine on pine sawflies. Oecologia<br />

WATT, and FORREST, G.I. 1991. The effect of previous defoliation of pole stage lodgepole pine on<br />

plant the growth and survival of pine beauty moth (Panolisflammea) larvae. Oecologia 86:31-35.<br />

20<br />

.... _!i5_ ¸i/_


STAND AND LANDSCAPE DIVERSITY AS A MECHANISM OF FOREST<br />

RESISTANCE TO INSECTS<br />

T. D. SCHOWALTER<br />

Department of Entomology, Oregon State University, C<strong>or</strong>vallis, Oregon 97331-2907, USA<br />

INTRODUCTION<br />

Plant vulnerability to herbiv<strong>or</strong>ous insects depends on both suitability (determined by nutritional and defensive<br />

fact<strong>or</strong>s) and exposure (determined by plant location relative to herbiv<strong>or</strong>e population sources). The imp<strong>or</strong>tance of plant<br />

biochemical defenses against herbiv<strong>or</strong>es has received much attention (Rosenthal and Janzen 1979, Cates and Alexander 1982,<br />

Harb<strong>or</strong>ne 1982, Coley et al. 1985). Herbiv<strong>or</strong>e populations typically are aggregated on particular trees that often differ from<br />

their neighb<strong>or</strong>s in chemical composition <strong>or</strong> other fact<strong>or</strong>s that indicate greater susceptibility to herbiv<strong>or</strong>es (Alstad and<br />

Edmonds 1983, L<strong>or</strong>io 1993). Plant compounds are imp<strong>or</strong>tant as feeding deterrents <strong>or</strong> toxins f<strong>or</strong> herbiv<strong>or</strong>es unable to adapt<br />

appropriate avoidance <strong>or</strong> detoxification mechanisms. However, herbiv<strong>or</strong>es have adapted various strategies f<strong>or</strong> avoiding <strong>or</strong><br />

detoxifying the chemical defenses of their hosts (Bernays and Woodhead 1982, McCullough and Wagner 1993).<br />

Variation in suitability among plants may be imp<strong>or</strong>tant both f<strong>or</strong> reducing selection f<strong>or</strong> herbiv<strong>or</strong>e adaptations and f<strong>or</strong><br />

minimizing host exposure to adapted herbiv<strong>or</strong>es. Herbiv<strong>or</strong>es seeking new hosts must both distinguish and be able to reach<br />

suitable hosts. Theref<strong>or</strong>e, plant suitability, apparency, and distance from herbiv<strong>or</strong>e populations function interactively to<br />

determine colonization by herbiv<strong>or</strong>es (Courtney 1986, Schowalter and Stein 11987).<br />

My objective in <strong>this</strong> paper is to evaluate plant diversity as a means of reducing host exposure to herbiv<strong>or</strong>es. I will<br />

consider diversity at the level of genetic variability within a particular species, and at the level of stand and landscape<br />

variability in community composition. These three levels constitute a nested hierarchy of diversity that can limit population<br />

growth of herbiv<strong>or</strong>es within stands and across landscapes.<br />

Genetic Diversity Within Species<br />

Populations of plants typically vary in genotype and susceptibility to herbiv<strong>or</strong>es. Herbiv<strong>or</strong>es usually prefer, and show<br />

highest survival and reproduction on, plants <strong>or</strong> plant species whose defenses can be tolerated <strong>or</strong> detoxified, given that these<br />

plants can be discovered (Courtney 1986). Most herbiv<strong>or</strong>es are specialists on relatively few (usually related) plant species,<br />

because avoidance behavi<strong>or</strong>s and detoxification mechanisms require an appropriate genetic template and are energetically<br />

expensive. Insects that feed on several plant species with distinct resistance mechanisms are subject to different selective<br />

f<strong>or</strong>ces on each host and may develop sibling species as host-specific demes diverge (Alstad and Edmonds 1983; Via 1990,<br />

1991; Bright 1993). Conversely, stands of plants with similar resistance mechanisms permit rapid adaptation by herbiv<strong>or</strong>es.<br />

Genetic diversity within a plant species confers resistance to herbiv<strong>or</strong>es by presenting a matrix of susceptible and<br />

non-susceptible hosts, but genetic effects are difficult to evaluate in natural f<strong>or</strong>ests where genotypic effects can be confounded<br />

by plant proximity to herbiv<strong>or</strong>e population sources <strong>or</strong> other fact<strong>or</strong>s varying geographically. Conifer seed <strong>or</strong>chards<br />

and nurseries provide an opp<strong>or</strong>tunity to examine the imp<strong>or</strong>tance of plant genotype independent of plant location. Herbiv<strong>or</strong>e<br />

impact on different genotypes replicated throughout a seed <strong>or</strong>chard <strong>or</strong> nursery can be compared to assess the extent to which<br />

injury is related to genotype.<br />

Schowalter and Haverty (1989) examined seed losses to two insect species, Contarinia <strong>or</strong>egonensis and Megastigmus<br />

spermotrophus, in a Pseudotsuga menziesii clonal seed <strong>or</strong>chard and in a progeny plantation in western Oregon. The various<br />

clones <strong>or</strong> families showed different degrees of resistance to the two insects (Table 1). Resistance to one insect species was<br />

Mattson, W.J., NiemelS., R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

21


Table 1.--Percentage seed lost to two cone and seed insect species among offspring of selected parental crosses in a<br />

Pseudotsuga menziesii progeny plantation in western Oregon. Means above the diagonal are seed losses to a midge,<br />

Contarinia <strong>or</strong>egonensis; values below the diagonal are losses to a chalcid, Megastigmus spermotrophus.<br />

x Parent 1 2 3 4 5 6 7 8 9 10 11 12 Midge<br />

Parent Mean _<br />

1 u 51 43 55 56 54 68 39 43 67 53<br />

2 8 -- 58 55 84 75 62 85 76 51 59 67<br />

3 10 -- 54 61 79 65 60 59 58 62<br />

4 13 17 6 u 59 63 57 52 75 55 52 62 57<br />

5 6 7 11 -- 69 62 55 63 68 40 62<br />

6 10 10 24 6 -- 59 56 60 58 61 67 66<br />

7 6 6 8 7 16 6 -- 59<br />

8 4 18 13 7 10 m 65<br />

9 5 14 13 8 14 _ 66<br />

10 2 11 6 11 4 20 -- 59<br />

11 9 12 15 16 10 22 -- 51<br />

12 6 5 2 15 19 _ 63<br />

Chalcid<br />

Mean _ 7 9 12 14 8 14 8 10 11 9 14 9<br />

195% CI = 3.3-4.1 f<strong>or</strong> the midge and 1.2-1.4 f<strong>or</strong> the chalcid<br />

22 i<br />

ii


not related to resistance to the second species. Of ten clones in the seed <strong>or</strong>chard that deviated significantly (based on 95%<br />

CI) from mean seed loss to either species, eight that were resistant to one species showed no resistance to the other. Four of<br />

these clones were highly susceptible to the second species. Only two clones were resistant <strong>or</strong> susceptible to both insects.<br />

Similar results were found in the progeny plantation (Schowalter and Haverty 1989). Parental crosses that were<br />

resistant to one insect generally were susceptible to the other. In <strong>this</strong> case, resistance appeared to be heritable as a dominant<br />

trait, based on the generally low seed losses f<strong>or</strong> progeny of resistant and susceptible parents (Table 1).<br />

Schowalter and Stein (1987) compared the extent of Lygus hesperus feeding on different seed sources (representing<br />

different genetic backgrounds) in a conifer seedling nursery. This insect is a mobile species that feeds primarily on agricultural<br />

crops :in a hit-and-run manner. Results indicated significant separate and interactive effects of conifer seed source<br />

(genotype) and of proximity to Lygus population sources in adjacent agricultural crops. The effect of plant proximity to<br />

herbiv<strong>or</strong>e populations, even at <strong>this</strong> small scale f<strong>or</strong> a relatively mobile herbiv<strong>or</strong>e, is surprising.<br />

Data from these studies demonstrate that, within monocultures of plants, a diversity of genotypes can limit resource<br />

availability and suitability f<strong>or</strong> particular herbiv<strong>or</strong>e species. This diversity represents a species-level defense that limits initial<br />

herbiv<strong>or</strong>e population growth. Genetic diversity within a monoculture is not sufficient to prevent herbiv<strong>or</strong>e outbreaks over<br />

long time periods, especially when conditions that stress plants and/<strong>or</strong> inhibit production of plant defenses permit herbiv<strong>or</strong>e<br />

population growth (Waring and Pitman 1983). When environmental conditions increase susceptibility of a given plant<br />

species, exposure to herbiv<strong>or</strong>es can be reduced by surrounding non-host plants and stands.<br />

Plant Species Diversity Within Stands<br />

Our view of plant species interactions traditionally has focused on competitive interactions. This view has supp<strong>or</strong>ted<br />

the tree farm (monoculture) approach to f<strong>or</strong>estry. Recent studies, however, are indicating that plant species interactions are<br />

m<strong>or</strong>e complex. Plant species often share myc<strong>or</strong>rhizae, contribute collectively to soil fertility through differential nutrient<br />

uptake and concentration in litter and rhizosphere, and increase the chemical complexity of the f<strong>or</strong>est aerosol (Visser :1986,<br />

Hunter and Arssen 1988). These mutualistic aspects of plant species interactions may reduce the likelihood of plant stress<br />

and apparency to herbiv<strong>or</strong>es, at least f<strong>or</strong> some combinations of plant species.<br />

If genetic diversity within monocultures can affect herbiv<strong>or</strong>es, then a diversity of plant species within a community<br />

matrix should limit herbiv<strong>or</strong>y to a greater extent. Studies with several insect species in different vegetation types have<br />

demonstrated that diverse vegetation limits overall herbiv<strong>or</strong>y. F<strong>or</strong> example, Root (1973), Kareiva (1983) and Turchin (1988)<br />

rep<strong>or</strong>ted that intermixed crops were subjected to lower levels of herbiv<strong>or</strong>y by insects than were monocultures of the same<br />

crops. Examples from f<strong>or</strong>ests are rare, largely because manipulating tree diversity f<strong>or</strong> experimental purposes is difficult, and<br />

natural variation in diversity is confounded by geographic variation in soils, aspect, and other fact<strong>or</strong>s that also affect herbiv<strong>or</strong>y<br />

(Schowalter and Filip 1993).<br />

Gara and Coster (1968), Johnson and Coster (1978), and Schowalter et al. (198 lb) rep<strong>or</strong>ted that southern pine beetle,<br />

Dendroctonusfrontalis, populations appeared to be sensitive to Pinus spp. density. Populations generally grew rapidly in<br />

dense pine stands and declined in sparse pine stands. However, Johnson and Coster (1978) and Schowalter et al. (1981b)<br />

rep<strong>or</strong>ted that large populations (> 100,000 beetles) developing under fav<strong>or</strong>able conditions also were capable of sufficient<br />

aggregation to colonize sparse hosts. Thinning experiments in western N<strong>or</strong>th America indicated that tree spacing also is<br />

critical to mountain pine beetle, Dendroctonus ponderosae, populations (Sartwell and Stevens 1975, Mitchell et al. 1983).<br />

The effects of tree species diversity were unclear; either hardwoods compete with pines and aggravate stress-related beetle<br />

activity (Hicks 1980) <strong>or</strong> hardwoods interfere with discovery of hosts (Belanger and Malac 1980).<br />

Schowalter and Turchin (1993) conducted a relatively unique experiment to test effects of Pinus spp. density and<br />

stand diversity on D. frontalis populations. They manipulated pine and hardwood basal areas in a :2x 2 fact<strong>or</strong>ial experiment<br />

and introduced equivalent D. frontalis populations into replicated plots to prevent confounding effects of plot proximity to<br />

beetle population sources. Infestations subsequently developed only in the dense pure pine stands and achieved infestation<br />

sizes (>10 dead trees) sufficient to warrant suppression in <strong>this</strong> treatment (Fig. 1). Infestations did not develop in dense pine<br />

stands where hardwoods were present <strong>or</strong> in low density stands, indicating that tree spacing and stand diversity both function<br />

to reduce insect population growth.<br />

23


uaa ,,.4 --" _" - LP/LH p<br />

ua_t _ 4 -''_''LP/HH<br />

_ a<br />

I--<br />

ua tu 3 _ ---'c> HP/LH _ r<br />

.!_ :. :--_::_---<br />

...:.. -.:..-. "-----'..:.- ----:-"------<br />

I 2 3 4<br />

TIME(Months)<br />

Figure 1.mCumulative pine m<strong>or</strong>tality to Dendroctonusfrontalis by pine and hardwood basal area treatments in Mississippi<br />

and Louisiana during 1989 and 1990. Vertical lines represent 1 SEM. Low pine (LP) = 11-14 m2/ha ba; high pine<br />

(HP) = 23-29 ba; low hardwood (LH) = 0-4 ba; high hardwood (HH) = 9-14 ba. Two infested trees were introduced<br />

into each experimental stand at the beginning of study. N-8.<br />

Schowalter (unpubl. data) compared arthropod abundances among replicated and intermixed plots representing young<br />

Douglas-fir, Pseudotsuga menziesii, plantations (10-15 years old), mature natural P. menziesii monocultures (100-150 years<br />

old), intact old-growth P. menziesii/Tsuga heterophylla (450 years, old) and old-growth P. menziesii shelterwood (450 years<br />

old) treatments in western Oregon. Western spruce budw<strong>or</strong>m, Ch<strong>or</strong>istoneura occidentalis, was significantly m<strong>or</strong>e abundant<br />

and caused twofold m<strong>or</strong>e defoliation in the mature monoculture than in other treatments. This difference also could reflect<br />

the greater predat<strong>or</strong> diversity and abundance in old-growth stands (Perry 1988, Schowalter 1989, T<strong>or</strong>gersen et al. 1990), <strong>or</strong><br />

unmeasured differences in mutualistic endophyte diversity <strong>or</strong> foliage quality among age classes (McCutcheon and Carroll<br />

1993).<br />

Stand Diversity Across Landscapes<br />

Diversity at the landscape level augments diversity at the stand level. As the diversity of stands representing different<br />

species composition <strong>or</strong> age classes increases, the distance between stands containing suitable resources increases (Perry I988,<br />

Schowalter 1989). Although herbiv<strong>or</strong>es are capable of dispersing over considerable distances, several fact<strong>or</strong>s reduce the<br />

likelihood of distant hosts being discovered <strong>or</strong> colonized. First, insect ability to perceive hosts over long distances is limited<br />

f<strong>or</strong> most species. Plant cues <strong>or</strong> other fact<strong>or</strong>s conveying host location become dissipated in the f<strong>or</strong>est aerosol, making distant<br />

hosts and hosts in diverse stands less apparent (Visser 1986). Second, survival decreases with distance as a result of longer<br />

exposure to m<strong>or</strong>tality agents and exhaustion of energy reserves, reducing insect ability to reach distant hosts. Theref<strong>or</strong>e,<br />

diverse landscapes should prevent localized outbreaks from spreading to distant hosts.<br />

Conversely, herbiv<strong>or</strong>e populations are promoted in landscapes that provide greater homogeneity of resources and few<br />

barriers to population spread. Maj<strong>or</strong> outbreaks typically occur in relatively homogeneous landscapes.<br />

Declining f<strong>or</strong>est health in eastern Oregon and Washington is largely the result of change in landscape diversity.<br />

F<strong>or</strong>ests in <strong>this</strong> region <strong>or</strong>iginally were a diverse matrix in which stands of shade-tolerant mixed-conifer Pseudotsuga/Abies/<br />

Pinus f<strong>or</strong>est in moist sites at higher elevations, n<strong>or</strong>th aspects, and riparian areas were embedded within an arid landscape<br />

dominated by sparse fire-tolerant Pinus/Larix f<strong>or</strong>est and savannah. Fir defoliat<strong>or</strong>s such as C. occidentalis occurred as local<br />

populations f<strong>or</strong>ced to search f<strong>or</strong> hosts aggregated within a largely inhospitable landscape. As a result of fire suppression and<br />

selective logging of Pinus and Larix over the past century, the fir f<strong>or</strong>est spread across <strong>this</strong> landscape. The landscape is now<br />

relatively homogeneous, dominated by f<strong>or</strong>ests of dense P. menziesii and Abies spp. Drought stress of these mesic species in<br />

addition to resource concentration has permitted C. occidentalis to reach epidemic population levels over most of <strong>this</strong> large<br />

area (Hadfield 1988).<br />

24<br />

i _


In contrast to the situation in eastern Oregon and Washington, C. occidentalis hist<strong>or</strong>ically has occurred at innocuous<br />

population levels in western Oregon and Washington, although localized outbreaks have occurred. Recent (and past) activity<br />

of C. occidentalis in western Oregon is concentrated around a maj<strong>or</strong> pass through the Cascade Range. This area has been<br />

affected by years of drought but also was accessible to epidemic C. occidentalis populations spilling over the pass from<br />

eastern Oregon (pets. obs.). Nevertheless, Schowalter (1989 and unpubl, data) found that, near <strong>this</strong> area, C. occidentalis was<br />

rare <strong>or</strong> absent in diverse old-growth f<strong>or</strong>ests but was abundant and causing measurable defoliation in mid-successional P.<br />

menziesii monocultures. Perry and Pitman (1983) compared suitability ofP. menziesii foliage from eastern and western<br />

Oregon f<strong>or</strong> C. occidentalis. They found that foliage from western Oregon was m<strong>or</strong>e susceptible to budw<strong>or</strong>m feeding and<br />

suggested that the greater diversity of trees, predat<strong>or</strong>s, and parasites in western Oregon has limited budw<strong>or</strong>m populations and<br />

minimized selection f<strong>or</strong> resistance to <strong>this</strong> insect. If diversity has been a maj<strong>or</strong> fact<strong>or</strong> preventing budw<strong>or</strong>m defoliation in<br />

western Oregon and Washington, then widespread commercial production of mid-successional P. menziesii f<strong>or</strong>ests may result<br />

in increasing C. occidentalis activity in <strong>this</strong> region.<br />

Outbreaks of D. frontalis in the southeastern U.S. also result from change in landscape diversity. Most of the land<br />

area in <strong>this</strong> region <strong>or</strong>iginally was vegetated by sparse woodlands and savannahs dominated by Pinus palustris, a species<br />

tolerant of the frequent fires and drought of <strong>this</strong> region and relatively resistant to bark beetles (Schowalter et al. 1981 a).<br />

Mesic riparian and bottomland f<strong>or</strong>ests included a diverse assemblage of intolerant species, including Pinus taeda and<br />

hardwoods. Bark beetles in <strong>this</strong> landscape would have been restricted primarily to scattered injured <strong>or</strong> diseased trees. Land<br />

conversion followed by eventuat abandonment and ref<strong>or</strong>estation led to establishment of dense stands of rapidly growing and<br />

commercially valuable P. taeda over most of <strong>this</strong> region. This species is susceptible to D. frontalis, resulting in devastating<br />

outbreaks across the region (Schowalter et al. 1981a, Schowalter and Turchin 1993).<br />

These observations indicate that diversity can be effective in limiting herbiv<strong>or</strong>e epidemiology. However, diversity<br />

: may not provide protection when herbiv<strong>or</strong>es reach large population size in a sufficient prop<strong>or</strong>tion of the surrounding landscape.<br />

SUMMARY<br />

Plant vulnerability to herbiv<strong>or</strong>es is a function of both biochemical suitability and exposure to herbiv<strong>or</strong>e populations.<br />

Plant exposure to herbiv<strong>or</strong>es can be minimized by plant diversity. Plant diversity at genotypic, species, and landscape levels<br />

presents herbiv<strong>or</strong>es with a mosaic of suitable and non-suitable hosts and non-hosts. This increases the distance herbiv<strong>or</strong>es<br />

must travel to reach new hosts and increases the difficulty of discovering suitable hosts hidden by non-hosts in diverse stands<br />

and landscapes. By contrast, limited genetic diversity within monocultures of commercially valuable trees provides a<br />

concentrated resource f<strong>or</strong> adapted herbiv<strong>or</strong>es. This resource inevitably will become m<strong>or</strong>e vulnerable to herbiv<strong>or</strong>es, especially<br />

during adverse conditions that stress plants and/<strong>or</strong> limit defensive capability. Localized outbreaks occur where isolated<br />

stands become vulnerable. These outbreaks can be restricted to isolated vulnerable stands in a diverse landscape that<br />

provides barriers to herbiv<strong>or</strong>e population spread. Regionwide outbreaks occur where a sufficient prop<strong>or</strong>tion of the landscape<br />

is occupied by suitable hosts. Large populations can expand into otherwise resistant stands and affect resistant <strong>or</strong> sparsely<br />

distributed hosts.<br />

LITERATURE CITED<br />

ALSTAD, D.N. and EDMONDS, G.F., Jr. 1983. Adaptation, host specificity, and gene flow in the black pineleaf scale, p.<br />

413-426. In Denno, R.F. and McClure, M.S., eds. Variable Plants and Herbiv<strong>or</strong>es in Natural and Managed Ecosystems.<br />

Academic Press, NY.<br />

AMMAN, G.D., MCGREGOR, M.D., SCHMITZ, R.F., and OAKES, R.D. 1988. Susceptibility of lodgepole pine to<br />

infestations by mountain pine beetles following partial cutting of stands. Can. J. F<strong>or</strong>. Res. 18: 688-695.<br />

BELANGER, R.P. and MALAC, B.E 1980. Silviculture can reduce losses from the southern pine beetle. Washington, DC:<br />

U.S. Department of Agriculture, Agric. Handb. 576.<br />

BERNAYS, E.A. and WOODHEAD, S. 1982. Plant phenols utilized as nutrients by a phytophagous insect. Science 216:<br />

201-202.<br />

25<br />

i


BRIGHT, D.E. 1993. Systematics of bark beetles, p. 23-36. In Schowalter, T.D. and Filip, G.M., eds. Beetle-Pathogen<br />

Interactions in Conifer F<strong>or</strong>ests. Academic Press, London.<br />

CATES, R.G. and ALEXANDER, H. 1982. Host resistance and susceptibility, p. 212-263. In Mitton, J.B. and Sturgeon,<br />

K.B., eds. Bark Beetles in N<strong>or</strong>th American Conifers: a System f<strong>or</strong> the Study of Evolutionary Biology. University of<br />

Texas Press, Austin.<br />

COLEY, ED., BRYANT, J.E, and CHAPIN, ES., III. 1985. Resource availability and plant antiherbiv<strong>or</strong>e defense. Science<br />

230: 895-899.<br />

COURTNEY, S.E 1986. The ecology of pierid butterflies: dynamics and interactions. Adv. Ecol. Res. 15:51-131.<br />

GARA, R.I. and COSTER, J.E. 1968. Studies on the attack behavi<strong>or</strong> of the southern pine beetle. III. Sequence of tree<br />

infestation within stands. Contrib. Boyce Thompson Inst. 24: 77-86.<br />

HADFIELD, J.S. 1988. Integrated pest management of a western spruce budw<strong>or</strong>m outbreak in the Pacific N<strong>or</strong>thwest.<br />

N<strong>or</strong>thw. Environ. J. 4: 301-312.<br />

HANSEN, E.M., GOHEEN, D.J., HESSBURG, EF., WITCOSKY, J.J., and SCHOWALTER, T.D. 1988. Biology and<br />

management of black-stain root disease in Douglas-fir, p. 63-80. In Harrington, T.C. and Cobb, F.W., Jr., eds.<br />

Leptographium Root Diseases on Conifers. American Phytophathological Society Press, St. Paul, MN.<br />

HARBORNE, J.B. 1982. Introduction to Ecological Biochemistry. Academic Press, London.<br />

HICKS, R.R., JR. 1980. Climate, site, and stand fact<strong>or</strong>s, p. 55-68. In Thatcher, R.C., Searcy, J.L., Coster, J.E., and Hertel,<br />

G.D., eds. The Southern Pine Beetle. Tech. Bull. 1631. Washington, DC: U.S. Department of Agriculture, F<strong>or</strong>est<br />

Service.<br />

HUNTER, A.E and ARSSEN, L.W. 1988. Plants helping plants. BioScience 38: 34-40.<br />

JOHNSON, EC. and COSTER, J.E. 1978. Probability of attack by southern pine beetle in relation to distance from an<br />

attractive host tree. F<strong>or</strong>. Sci. 24: 574-580.<br />

KAREIVA, E 1983. Influence of vegetation texture on herbiv<strong>or</strong>e populations: resource concentration and herbiv<strong>or</strong>e movemerit,<br />

p. 259-289. In Denno, R.E and McClure, M.S., eds. Variable Plants and Herbiv<strong>or</strong>es in Natural and Managed<br />

Systems. Academic Press, NY.<br />

LORIO, EL., Jr. 1993. Environmental stress and whole-tree physiology, p. 81-101. In Schowalter, T.D. and Filip, G.M.,<br />

eds. Beetle-Pathogen Interactions in Conifer F<strong>or</strong>ests. Academic Press, London.<br />

MCCULLOUGH, D.G. and WAGNER, M.R. 1993. Defusing host defenses: ovipositional adaptations of sawflies to plant<br />

resins, p. 157-172. In Wagner, M.R. and Raffa, K.E, eds. Sawfly Life Hist<strong>or</strong>y Adaptations to Woody Plants. Academic<br />

Press, San Diego, CA.<br />

MCCUTCHEON, T.L. and CARROLL, G.C. 1993. Genotypic diversity in populations of a fungal endophyte from Douglasfir.<br />

Mycologia 85: 180-186.<br />

MITCHELL, R.G., WARING, R.H., and PITMAN, G.B. 1983. Thinning lodgepole pine increases tree vig<strong>or</strong> and resistanc_<br />

to mountain pine beetle. F<strong>or</strong>. Sci. 29: 204-211.<br />

PERRY, D.A. t988. Landscape pattern and f<strong>or</strong>est pests. N<strong>or</strong>thwest. Environ. J. 4: 213-228.<br />

PERRY, D.A. and PITMAN, G.B. 1983. Genetic and environmental influences in host resistance to herbiv<strong>or</strong>y: Douglas-fir<br />

and the western spruce budw<strong>or</strong>m. Z. ang. Entomol. 96: 217-228.<br />

26


ROOT, R.B. 1973. Organization of a plant-arthropod association in simple and diverse habitats: the fauna of collards<br />

(Brassica oleracea). Ecol. Monogr. 43" 995-124.<br />

ROSENTHAL, G.A. and JANZEN, D., eds. 1979. Herbiv<strong>or</strong>es: Their Interaction with Secondary Plant Metabolites. Academic<br />

Press, NY.<br />

SARTWELL, C. and STEVENS, R.E. 1975. Mountain pine beetle in ponderosa pine: prospects f<strong>or</strong> silvicultural control in<br />

second-growth stands. J. F<strong>or</strong>. 73: 136-140.<br />

SCHOWALTER, T.D. 1989. Canopy arthropod community structure and herbiv<strong>or</strong>y in old-growth and regenerating f<strong>or</strong>ests in<br />

western Oregon. Can. J. F<strong>or</strong>. Res. 19: 318-322.<br />

SCHOWALTER, T.D. and FILIR G.M., eds. 1993. Beetle-Pathogen Interactions in Conifer F<strong>or</strong>ests. Academic Press,<br />

London.<br />

SCHOWALTER, T.D. and HAVERTY, M.I. 1989. Influence of host genotype on Douglas-fir seed losses to ContarMia<br />

<strong>or</strong>egonensis (Diptera: Cecidomyiidae) and Megasti'gmus spermotrophus (Hymenoptera: T<strong>or</strong>ymidae) in western<br />

Oregon. Environ. Entomol. 18: 94-97.<br />

SCHOWALTER, T.D, and STEIN, J.D. 1987. Influence of Douglas-fir seedling provenance and proximity to insect population<br />

sources on susceptibility to Lygus hesperus (Heteroptera: Miridae) in a f<strong>or</strong>est nursery in western Oregon.<br />

Environ. Entomol. 16: 984-986.<br />

SCHOWALTER, T.D. and TURCHIN, R 1993. Southern pine beetle infestation development: interaction between pine and<br />

hardwood basal areas. F<strong>or</strong>. Sci. 39: 201-210.<br />

SCHOWALTER, T.D., COULSON, R.N., and CROSSLEY, D.A., Jr. 198la. Role of southern pine beetle and fire in<br />

maintenance of structure and function of the southeastern coniferous f<strong>or</strong>est. Environ. Entomol. 10: 821-825.<br />

SCHOWALTER, T.D., POPE, D.N., COULSON, R.N., and FARGO, W.S. 198 lb. Patterns of southern pine beetle<br />

(Dendroctonusfrontalis Zimm.) infestation enlargement. F<strong>or</strong>. Sci. 27: 837-849.<br />

TORGERSEN, T.R., MASON, R.R., and CAMPBELL, R.W. 1990. Predation by birds and ants on two f<strong>or</strong>est insect pests in<br />

the Pacific N<strong>or</strong>thwest. In M<strong>or</strong>rison, M.L., Ralph, C.J., Verner, J., and Jehl, J.R., Jr., eds. Avian F<strong>or</strong>aging: The<strong>or</strong>y,<br />

Methodology, and Applications. Cooper Ornithological Society, Studies in Avian Biology 13: 14-19.<br />

TURCHIN, R 1988. The effect of host-plant density on the numbers of Mexican bean beetles, Epilachna varivestis. Am.<br />

Midl. Nat. 119: 15-20.<br />

VIA, S. 1990. Ecological genetics and host adaptation in herbiv<strong>or</strong>ous insects: the experimental study of evolution in natural<br />

and agricultural systems. Annu. Rev. Entomol. 35: 421-446.<br />

VIA, S. 1991. The genetic structure of host plant adaptation in a spatial framew<strong>or</strong>k: demographic variability among reciprocally<br />

transplanted pea aphid clones. Evolution 45: 827-852.<br />

VISSER, J.H. 1986. Host od<strong>or</strong> perception in phytophagous insects. Annu. Rev. Entomol. 31 : 12I- 144.<br />

WARING, R.H. and PITMAN, G.B. 1983. Physiological stress in lodgepole pine as a precurs<strong>or</strong> f<strong>or</strong> mountain pine beetle<br />

attack. Z. ang. Entomol. 96: 265-270.<br />

WITCOSKY, J.J., SCHOWALTER, T.D., and HANSEN, E.M. 1986a. The influence of precommercial thinning on the<br />

colonization of Douglas-fir by three species of root-colonizing insects. Can. J. F<strong>or</strong>. Res. 16: 745-749.<br />

WITCOSKY, J.J., SCHOWALTER, T.D., and HANSEN, E.M. 1986b. Hylastes nigrinus (Coleoptera: Scolytidae), Pissodes<br />

fasciatus, and Steremnius carinatus (Coleoptera: Curculionidae) as vect<strong>or</strong>s of black-stain root disease of Douglas-fir.<br />

Environ. Entomol. 15: 1090-1095. 27


' T<br />

ON NEIGHBOR EI_FEC S IN PLANT-HERBIVORE INTERACTIONS w e<br />

g<br />

JUHA TUOMI, MAGNUS AUGNER, and PATRIC NILSSON<br />

Department of Ecology, The<strong>or</strong>etical Ecology, Lund University, T Ecology Bldg., S-223 62 Lund, Sweden<br />

INTRODUCTION<br />

Coevotution involves genetic changes that occur in both plant and herbiv<strong>or</strong>e populations over evolutionary time.<br />

Genetic changes in turn depend on fitnesses associated with given plant and herbiv<strong>or</strong>e genotypes. This leads us to the<br />

ecologicaI time scale because fitness is determined by functional interactions between individual plant units and herbiv<strong>or</strong>es<br />

over their life-spans. Such interactions can be f<strong>or</strong>malized as a dependency of fitness on plant and herbiv<strong>or</strong>e phenotypes in a<br />

giver_context of interaction,<br />

The basic features of co-evolutionary processes depend primarily on the direct effects that an individual's phenotype<br />

has of_its fitness. However, as pointed out by Atsatt and O' Dowd (1976) and Rhoades (1979) among others, the actual<br />

interactions are often much m<strong>or</strong>e complicated due to various indirect effects when fitness depends on what the neighb<strong>or</strong>s are<br />

doing. We label these indirect relations as neighb<strong>or</strong> effects (Eshel 1972), and argue that they may have a fundamental<br />

.....<br />

imp<strong>or</strong>tance in plant-herbiv<strong>or</strong>e interactions (Mattson et al. 1991).<br />

We discuss two aspects of indirect interactions between plants. First, in some cases palatable plants gain a benefit by<br />

growing close to unpalatable plants. This possibility was first discussed by Tahvanainen and Root (1972) and Atsatt and<br />

O'Dowd (1976), and may be called "associational refuge" effect (Pfister and Hay 1988). Several experiments have demonstrated<br />

a refuge effect (McNaughton 1978, Hay 1986), but not always (McNaughton 1978, Danell et al. 1991, ttj;,ilten et al.<br />

1993).<br />

Second, it is frequently assumed that herbiv<strong>or</strong>es can choose their host plants depending on their nutritive value and<br />

toxicity. This presumes explicitly that herbiv<strong>or</strong>es can recognize the host plant and "read" cues that c<strong>or</strong>relate with palatability.<br />

If the herbiv<strong>or</strong>e can generalize the cues (Launchbaugh and Provenza 1993), unpalatable plants that repel herbiv<strong>or</strong>es will<br />

naturally gain a benefit themselves. Likewise, the herbiv<strong>or</strong>e should avoid other plants with these very same cues. That is<br />

'_ (3" "<br />

why we expect that plant defenses may imply syner_:st_c benefits" in the sense that a plant implies a benefit f<strong>or</strong> other<br />

individuals sharing the same phenotype (Nee 1989,Guilf<strong>or</strong>d and Cuthill 1991, Rosenberg 199 l, and Tuomi and Augner<br />

993), .....<br />

We propose that these neighb<strong>or</strong> effects may well shape some aspects of plant defenses. F<strong>or</strong> <strong>this</strong> purpose, we present a<br />

simplified model system where plants are :interactingin pairs. Then we analyze the model system in a game the<strong>or</strong>etical<br />

"<br />

context by assuming three phenotypes: non-defensive type (N), and two defensive types - one which kills the invading<br />

herbiv<strong>or</strong>es (D), and another that provides herbiv<strong>or</strong>es a signal over which to generalize (S).<br />

MODEL SYSTEM<br />

We assume a plant population where individual plants are interacting in pairs, i.e., in trait groups (Wilson 1975) of<br />

two plants. Each pair has a risk of herbiv<strong>or</strong>y, say m: that is a function of grazing intensity over the whole population and of<br />

the phenotypic composition of the given trait group. We call the plants the player and the neighb<strong>or</strong> (<strong>or</strong> the opponent). An<br />

invading herbiv<strong>or</strong>e selects the player with probability s and the neighb<strong>or</strong> with probability 1-s. We will, f<strong>or</strong> simplicity, assume<br />

that the consequent risk of herbiv<strong>or</strong>y is equal f<strong>or</strong> both the player and the neighb<strong>or</strong> so that mps = mp(1-s) = m in all trait<br />

groups, in other w<strong>or</strong>ds, invading herbiv<strong>or</strong>es can select neither between trait-groups n<strong>or</strong> between the player and the neighb<strong>or</strong>.<br />

Mattso__,,W.J., Niemela, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Secy. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

28


We tlhus neglect two potential sources of neighb<strong>or</strong> effects because experienced herbiv<strong>or</strong>es may well be able to evaluate the<br />

expected value of a patch <strong>or</strong> a plant from a distance, e.g., by appearance <strong>or</strong> scent. Hjfilten et al. (1993) have discussed in a<br />

greater length how selection between and within patches can affect associational refuges.<br />

In sh<strong>or</strong>t, our herbiv<strong>or</strong>es are naive, but we assume that they can learn to evaluate the defensive status of the plants if<br />

appropriate cues are present. The consumption process is divided into two parts (bites). From our earlier assumptions, it<br />

follows that the risk of the first bite is m. Each bite by a herbiv<strong>or</strong>e implies a fitness cost f<strong>or</strong> the plant. Plant defenses can in<br />

various ways reduce the total cost of herbiv<strong>or</strong>y: reducing the fitness cost caused by a herbiv<strong>or</strong>e per bite, increasing m<strong>or</strong>tality<br />

(d) of herbiv<strong>or</strong>es during a bite, and inducing feeding aversion that can be either unconditional (the herbiv<strong>or</strong>e does not take the<br />

second bite of any food) <strong>or</strong> conditional (the herbiv<strong>or</strong>e does not take the second bite if a specific cue is present). In our case of<br />

two bites, unconditional feeding aversion is equivalent to increased m<strong>or</strong>tality as in both cases further herbiv<strong>or</strong>y is stopped.<br />

We thus consider only conditional feeding aversion and the parameter a indicates the probability that feeding aversion is<br />

induced during a bite. We also assume that the cost of herbiv<strong>or</strong>y is h if the herbiv<strong>or</strong>e survives and no feeding aversion is<br />

induced, while the cost of herbiv<strong>or</strong>y will be h' if the herbiv<strong>or</strong>e happens to die <strong>or</strong> if a feeding aversion is induced during a bite<br />

(O < h' < h). We assume specifically that non-defensive (N) plants cause no m<strong>or</strong>tality (d - 0, a - 0), while there are two kinds<br />

of defensive plants; those (D) that will kill the herbiv<strong>or</strong>e during the bite (d = 1, a - 0), and those (S) that cause a conditional<br />

feeding aversion bef<strong>or</strong>e any toxic effects appear on herbiv<strong>or</strong>e survival (a - 1, d = 0).<br />

Both kinds of defenses (D and S) reduce the average costs that herbiv<strong>or</strong>y implies to the player. In addition to these<br />

direct fitness effects, various neighb<strong>or</strong> effects can arise if herbiv<strong>or</strong>es are allowed to move from the player to the neighb<strong>or</strong> and<br />

vice versa. We assume that a surviving herbiv<strong>or</strong>e moves after the first bite with a probability e. F<strong>or</strong> simplicity, feeding<br />

aversion is not allowed to affect <strong>this</strong> probability, and the parameters h, h', d, and a assume the same values during the second<br />

bite as earlier during the first bite. Taken together, these assumptions imply that the average cost (H) of herbiv<strong>or</strong>y over two<br />

bites will be<br />

H=[(1-di)(1-ai)h + (1-di)ah' + d'ih](1 +nii+nij)<br />

ni_= (1-d,)(1-a)(1-e)<br />

nij = { (1-di)(1-ai)ei (i,j=S) }<br />

(1-dj)ej (i v j aS)<br />

where n_ implies the probability of a second bite f<strong>or</strong> the player by a herbiv<strong>or</strong>e initially invading the player (denoted by i), and<br />

n_j by a herbiv<strong>or</strong>e moving from the neighb<strong>or</strong> (denoted by j) to the player. The latter term will be (1-d)(1-a)ej if feeding<br />

aversion induced by the neighb<strong>or</strong> protects the player (i.e., both S), and (1-d)e if the neighb<strong>or</strong> does not induce a feeding<br />

aversion (i.e., the neighb<strong>or</strong> N <strong>or</strong> D) <strong>or</strong> if the player does not possess the cue required f<strong>or</strong> the maintenance of feeding aversion<br />

(i.e., the neighb<strong>or</strong> S and the player either N <strong>or</strong> D). In our specific case (ei = e. l = e), H = h[1 + (l-e) + n_j]f<strong>or</strong> a player adopting<br />

N (d = 0, a = 0), and H = h'(1 + nij)f<strong>or</strong> a player adopting D (d = 1, a = 0) <strong>or</strong> S (d = 0, a = 1). When H is multiplied by m we<br />

will get, following Augner et al. (1991), a measure of the herbiv<strong>or</strong>y load, Hm, f<strong>or</strong> the payoff matrix.<br />

This highly simplified model system elucidates clearly that the fitness of the player may well depend on the neighb<strong>or</strong>.<br />

In our case, <strong>this</strong> is so if herbiv<strong>or</strong>es can move between the plants (0 < e


¢D<br />

Neighb<strong>or</strong><br />

N D S<br />

N -2m -(2-e)m -2m<br />

>_ D 1 (l+e)m-C 1 m-C 1<br />

¢_ -_ - _ - _ (l+e)m-C<br />

, ,,<br />

S -_1 (l+e)m-C(l+k) -21 m-C(l+k) -21 m-C(l+k)<br />

Figure 1.--Payoff matrix f<strong>or</strong> a game between two plants showing the payoffs to the player adopting either nondefensive<br />

strategy (N), lethal defense (D), <strong>or</strong> defense inducing conditional feeding aversion (S). (m) denotes the risk of<br />

herbiv<strong>or</strong>y, C as well as k indicate the costs of defenses, and e is the probability that a herbiv<strong>or</strong>e moves from one plant<br />

to another.<br />

the model so that D and S types have similar payoffs when playing against N; the only difference is the cost of defense. Our<br />

primary interest lies in the qualitative differences between the two defensive types D (lethal defense) and S (conditional<br />

feeding aversion). We define parameter space where these phenotypes are evolutionarily stable strategies (ESS) in the three<br />

subgames (Figs. 2-4). Finally, we study the dynamics of the game when all three strategies are taken account (Figs. 5 and 6).<br />

In all cases, we have random associations and the population frequencies of N, D, and S are q, p, and r respectively.<br />

A Disadvantage of Lethal Defense<br />

The first subgame of N and D contrasts a nondefensive strategy (N) against a defensive strategy (D) that kills all<br />

invading herbiv<strong>or</strong>es and that implies a cost, C. When the trait groups are randomly derived from the population frequencies q<br />

= l - p f<strong>or</strong> N and p f<strong>or</strong> D, the fitnesses will be<br />

W = Wo - (2-pe) m<br />

W_ =W 0- C- 1/2 [1 + (1-p)e]m<br />

where W o is a basic level of fitness. When e = 0, selection will exclusively fav<strong>or</strong> either N <strong>or</strong> D depending on the balance<br />

between the cost of defense (C) and the risk of herbiv<strong>or</strong>y (m). When e > 0, some herbiv<strong>or</strong>es will move from N to D. This<br />

wifl reduce the herbiv<strong>or</strong>y load of N and increase that of D in mixed pairs. As a consequence, we will have three possibilities<br />

f<strong>or</strong> different combinations of C/m and e (Fig. 2a):<br />

(1) N will be an ESS f<strong>or</strong> C/m> (3 - e)/2.<br />

(2) D will be ESS tbr C/m < (3 - 2e)/2.<br />

(3) Stable polym<strong>or</strong>phism (0 < _ < 1) is obtained when neither N n<strong>or</strong> D are ESSs.<br />

3O


c__ c__<br />

1T1 Ill<br />

a) a)<br />

0 0<br />

0 e 1 o e 1<br />

W N<br />

Z Z<br />

i<br />

I<br />

b) b)<br />

! I I<br />

0 _ p 1 0 9 r 1<br />

Figure 2.--ESS-conditions in N-D subgame contrasting Figure 3.--ESS-conditions in N-S subgame contrastnondefensive<br />

strategy (N) and lethal defense (D): ing nondefensive strategy (N) and defense (S)<br />

[a] The parameter areas where N and D are pure that induces conditional feeding aversion: [a]<br />

ESSs are shaded, while in the blank area neither The parameter areas where N and S are ESSs<br />

of them is an ESS; and [b] Fitnesses of N and D are shaded; and [b] Fitnesses of N and S as a<br />

as a function of population frequency (1_) of D function of the population frequency (r) of S<br />

when 0 < _ < 1. I_ = stable equilibrium; C, m, and when 0 < _ < 1. _ = unstable equilibrium; C,<br />

e as in Fig. 1. m, and e as in Fig. 1.<br />

31


Z<br />

C ,,,,, ,,,,<br />

m<br />

0<br />

a)<br />

0 e I<br />

I<br />

I<br />

I<br />

ws<br />

b)<br />

I<br />

0 _" r 1<br />

Figure 4.--ESS-conditions in D-S subgame contrasting lethal defense (D) and defense (S) that induces conditional<br />

feeding aversion: [a] The parameter areas where D and S are ESSs are shaded (k > 0); and [b] Fitnesses of D and<br />

S as a function of the population frequency (r) of S when 0 < } < 1. T 2= unstable equilibrium; C, m, and e as in<br />

Fig. 1.<br />

32


c_ >l.s S S S<br />

m<br />

N < DN DN < D<br />

c DN DN D<br />

k =0 k = 0.4- k = 1<br />

Figure 5.--Basic game dynamics when neighb<strong>or</strong> effects are excluded (e = 0). N, D, and S c<strong>or</strong>respond to populations where<br />

q = 1, p = 1 and r = 1 respectively.<br />

The third situation c<strong>or</strong>responds to "evolutionarily stable" associational refuges as both plant types coexists (Sabelis and de<br />

Jong 1988, Tuomi and Augner 1993). The coexistence is possible because both types have a selective advantage when rare,<br />

but a disadvantage when common (Fig. 2b). In our case, stable polym<strong>or</strong>phism requires that D has a higher fitness value f<strong>or</strong> p<br />

close to 0. However, when p increases toward 1, WN increases m<strong>or</strong>e steeply than W D. The slope of W Nas a function p is em,<br />

while the slope of W Dis (1/2)era. The reason f<strong>or</strong> <strong>this</strong> difference is that h' = 1/2. Consequently, the associational protection<br />

has a greater value f<strong>or</strong> N that suffers h = 1, while D suffers only a half of that value. If both N and D would win equally from<br />

the neighb<strong>or</strong>s' defenses (i.e., h' = h), associational protection would not maintain stable polym<strong>or</strong>phism in the present case.<br />

The maj<strong>or</strong> lesson of <strong>this</strong> subgame is that the associational effects counteract the evolution of lethal defenses. We<br />

have proposed that <strong>this</strong> implies a dilemma f<strong>or</strong> plant defenses: the defenses should protect the plant without benefiting<br />

nondefensive neighb<strong>or</strong>s too much (Tuomi et al. submitted). If the defensive plants kill all herbiv<strong>or</strong>es, the nondefensive type<br />

can easily invade. Consequently, one could expect that "less lethal" defenses could do a better job in the long run. However,<br />

when adopting sublethal defenses the plant itself will suffer m<strong>or</strong>e damage by herbiv<strong>or</strong>es. There are at least two solutions f<strong>or</strong><br />

the dilemma. First, if a plant is allowed to adopt a lethality-level (d) between 0 and 1, associational refuge effects select f<strong>or</strong><br />

sublethal defenses with d < 1 (Tuomi et al. submitted). Second, another solution of the dilemma could be a defense strategy<br />

(S) that is potentially "lethal" to the herbiv<strong>or</strong>es, but that is associated with chemical <strong>or</strong> visual traits which herbiv<strong>or</strong>es can use<br />

as cues in <strong>or</strong>der to avoid the potentially "lethal" defenses. Below, we analyze <strong>this</strong> second possibility. Since we have set d = 0<br />

f<strong>or</strong> S, we thus assume that feeding aversion is induced bef<strong>or</strong>e the amount of ingested food reaches any toxic level.<br />

33


C__=l.G S S S<br />

m<br />

N DNv _-. DN D<br />

c o8 S S S<br />

m- "<br />

N DN .. DN D<br />

P P P<br />

c 04 S S S<br />

DN l > -DN D<br />

k =0 k =0.4 k = 1<br />

Figure 6.--Basic game dynamics when herbiv<strong>or</strong>es are allowed to move between the interacting plants (e = 1). N, D, and<br />

S c<strong>or</strong>respond to populations where q = 1, p = I, and r = 1, respectively. Q is an unstable equilibrium and R is a<br />

saddle point, while P can be a saddle point <strong>or</strong> a stable equilibrium. Traject<strong>or</strong>ies will eventually approach the<br />

points denoted by the dots.<br />

Conditional Feeding Aversion<br />

We f<strong>or</strong>mulated the present model so that the player is subject to the exactly same herbiv<strong>or</strong>y load when adopting either<br />

D <strong>or</strong> S, and when the neighb<strong>or</strong> is N (Fig. 1). There are two maj<strong>or</strong> differences when N-S subgame is compared to N-D<br />

subgame. First, we assume that S and D may imply different costs f<strong>or</strong> the plant. We have defined that the cost of S is<br />

C(l+k). To be conservative, we assume that k is non-negative indicating that it is equally <strong>or</strong> m<strong>or</strong>e costly to induce and<br />

maintain feeding aversion than to kill the herbiv<strong>or</strong>e. If both D and S require an equal investment in defensive compounds,<br />

34


:hen k > 0 suggest that the cue <strong>or</strong> the signal associated with S is costly. Second, the maj<strong>or</strong> difference between S and D is that<br />

does not allow N to escape from herbiv<strong>or</strong>y in mixed pairs. That is why the fitness of N, when playing against S, reduces to<br />

whereas Ws will be<br />

W N= W0- 2m<br />

Ws = W0 - C(I + k)-l/2 [l+(1-r)e]m<br />

where r is the population frequency of S. The population frequency of N is q = 1 - r. Also in <strong>this</strong> case we have three<br />

possibilities (Fig. 3a):<br />

(1) N will be an ESS when C/m > (3 - e)/2(l+k).<br />

(2) S will be an ESS when C/m < 3/2(1+k).<br />

(3) There will be a unstable equilibrium 0 < _ < 1 when both N and S are ESSs.<br />

In other w<strong>or</strong>ds, no stable coexistence is possible since S will be an ESS above the equilibrium frequency ?, and N will be an<br />

ESS below the equilibrium (Fig. 3b).<br />

In summary, <strong>this</strong> subgame has two interesting aspects. The first is that nondefensive plants do not benefit from S<br />

because herbiv<strong>or</strong>es are assumed to be able to make a distinction between N and S types. The second is that the defensive<br />

plants will benefit from each other if feeding aversion is maintained by a cue that the neighb<strong>or</strong>s share. As a consequence, the<br />

fitness of defensive plants will increase as the frequency of S approaches 1, while the fitness of nondefensive plants remains<br />

constant (Fig. 3b). This situation c<strong>or</strong>responds to "synergistic selection" discussed by Guilf<strong>or</strong>d and Cuthill (1991) in the<br />

evolution of aposematism, and by Tuomi and Augner (1993) in relation to plant defenses.<br />

Costs of Defenses<br />

When allowing D and S to play against each other, a plant adopting D will obtain the same payoffs as when playing<br />

against N (Fig. 2). A plant adopting S will obtain the same payoffs when the neighb<strong>or</strong> is D <strong>or</strong> S. When r is the population<br />

frequency of S, we get the following fitnesses<br />

W D=W0 - C - 1/2 m(1 +er)<br />

W s =W 0- C(1 + k)- 1/2 m<br />

where W Ddeclines as a function of er, while Ws is independent of both e and r. There are four qualitatively different<br />

situations:<br />

(1) S will be a pure ESS either ifk < 0, <strong>or</strong> ifk = 0 and e > 0.<br />

(2) S and D are equally fit if k = 0 and e = 0.<br />

(3) D will be a pure ESS if k > 0 and C/m > e/2k.<br />

(4) There is an unstable equilibrium 0 < } < 1 if k > 0 and C/m < e/2k.<br />

Consequently, the costs of defenses play a prominent role here. If S is expensive relative to D, lethal defense is most<br />

likely to be selected f<strong>or</strong>. Fig. 4a indicates the last two situations obtained f<strong>or</strong> k > 0. The unstable equilibrium<br />

? = 2Ck/em<br />

represents the point where the fitness curves intersect each other (i.e., Ws = WD,Fig. 4b). When Fmoves closer to zero,<br />

the situation becomes m<strong>or</strong>e fav<strong>or</strong>able to S. This will happen when k becomes smaller, <strong>or</strong> when e and m become larger.<br />

The reverse changes will acc<strong>or</strong>dingly fav<strong>or</strong> D. We interpret <strong>this</strong> so that lethal defenses may be m<strong>or</strong>e economical against<br />

rare herbiv<strong>or</strong>es (m small) and sedentary herbiv<strong>or</strong>es (e small), while defenses inducing conditional feeding aversion<br />

should be m<strong>or</strong>e effective against common herbiv<strong>or</strong>es that are mobile and move from a plant to another (both m and e<br />

high).<br />

35


Game Dynamics<br />

When the above subgarnes are combined, there arises a number of possible dynamic solutions f<strong>or</strong> the game where a_t<br />

three strategies are taken account. We have chosen to describe some basic situations f<strong>or</strong> e = 0 (Fig. 5) and e = 1 (Fig. 6) whe_,<br />

k 2 0. The population frequencies of N, D and S are q, p and r respectively (q + p + r = 1), and the equilibria discussed abo'_e<br />

are denoted by P = (q, fi, O) (the stable equilibrium of the N-D subgame), Q = (_t,0, _) (the unstable equilibrium of the N-S<br />

subgame), and R = (0, _,'?) (the unstable equilibrium of the D-S subgame).<br />

When e = 0, them are no neighb<strong>or</strong> effects and hence no frequency dependency. Now the game dynamics can be<br />

directly derived from fitness ranks, as<br />

WN= W_ - 2m<br />

Ws =W_)- C(1 + k)- l/2m<br />

Ws = W0-C(1 +k)- l/2m<br />

When C is sufficiently low relative to m, selection will fav<strong>or</strong> a defensive strategy, either S <strong>or</strong> D. F<strong>or</strong> k > O, selection<br />

will always fav<strong>or</strong> D over S. On the other hand, when C is high relative to m, N will always be an ESS (Fig. 5).<br />

When e = l, the situation is a little bit m<strong>or</strong>e complicated as now fitnesses depend on population composition. Using<br />

the standard methods (Tayl<strong>or</strong> and Jonker 1978, Zeeman 1981,Bomze 1983, Hofbauer and Sigmund 1988), we can analyze<br />

the local stability of equilibria (Fig. 61).Also in <strong>this</strong> case, selection will fav<strong>or</strong> a defensive strategy if C is sufficiently low<br />

relative to m. However, now S will have an advantage over D when k = 0. F<strong>or</strong> a higher values of k (e.g., k = 0.4 <strong>or</strong> 1.0, Fig.<br />

6)_the equilibria Q and R appear with the consequence that, depending on the initial state, the population may evolve toward<br />

S, <strong>or</strong> alternatively to D (<strong>or</strong> P). Finally, ifC is high relative to m, INwill have an advantage over both D and S.<br />

This analysis confin,as the earlier result of the D-S subgame that high m and e fav<strong>or</strong> S (Fig. 6), and low e fav<strong>or</strong>s D<br />

(Fig. 5)_ However, the suggestion that low m fav<strong>or</strong>s D over S, does not imply that D is an ESS. When m is low enough, N<br />

can invade and either out-compete both defensive types <strong>or</strong> establish stable polym<strong>or</strong>phism with D.<br />

CONCLUSIONS<br />

Plant de:lenses have both direct effects on plant fitness and indirect effects that arise from interactions between<br />

neighb<strong>or</strong>ing plants. Although direct fitness effects obviously are the primary f<strong>or</strong>ces of selection, neighb<strong>or</strong> effects may also<br />

shape some basic aspects of plant defenses. This is especially so if fitness differences due to the direct effects of plant<br />

phe_otypes are marginal.<br />

We analyzed two types of neighb<strong>or</strong> effects when herbiv<strong>or</strong>es are allowed to move from one plant to another:<br />

(t) Palatable plants can benefit from the defenses of their unpalatable neighb<strong>or</strong>s. Defenses can, f<strong>or</strong> instance, increase<br />

m<strong>or</strong>tality among herbiv<strong>or</strong>es <strong>or</strong> induce :unconditional feeding aversion with the consequence that the herbiv<strong>or</strong>y load of<br />

palatable plants is low in mixed trait groups. Associational protection can maintain stable polym<strong>or</strong>phism if palatable plants<br />

gain m<strong>or</strong>e in fitness than unpalatable plants when growing close to an unpalatable neighb<strong>or</strong>. We thus expect that associational<br />

protection counteracts the evolution of plant defenses that do not provide herbiv<strong>or</strong>es any signal over which to generalize_<br />

(2) Unpalatable plants can selectively benefit other such plants if they share a trait that herbiv<strong>or</strong>es can use as a basis<br />

of their food intake and their choice f<strong>or</strong> host plants. In that case, palatable and unpalatable plants that do not share the signal<br />

do not gain assc_:iational protection. Defenses associated with a signal are most effective against abundant and mobile<br />

he:rbiv<strong>or</strong>es, and their advantage is greatest when they are common.<br />

Conseque__tly, we expect that there should be some c<strong>or</strong>respondence between the design of plant defenses and the<br />

sens<strong>or</strong>y and Iearning mechanisms of herbiv<strong>or</strong>es. Undoubtedly, th<strong>or</strong>ns and spines are defensive characters that can functi<strong>or</strong>l<br />

36


oth as protective weapons and as potential signals. If <strong>this</strong> also holds f<strong>or</strong> chemical defenses, neighb<strong>or</strong> effects might have a<br />

_ndamental imp<strong>or</strong>tance in shaping plant defenses. This may well be so as, because acc<strong>or</strong>ding to Provenza et al. (1992),<br />

mmmalian herbiv<strong>or</strong>es learn to select food items through two interrelated systems. The affective system integrates the taste<br />

f food and its post-ingestive consequences, while the cognitive system integrates the od<strong>or</strong> and sight of food and its taste.<br />

tecause they have to sample plants in <strong>or</strong>der to adjust food intake to avoid intoxication (Provenza et al. 1992), the expected<br />

erbiv<strong>or</strong>y load of a plant is likely to depend both on its own defenses and on the defensive status of its neighb<strong>or</strong>s.<br />

SUMMARY<br />

In <strong>or</strong>der to be evolutionary stable, plant defenses should not benefit palatable neighb<strong>or</strong>s too much. If they do,<br />

_alatable plants can invade and eventually either out-compete unpalatable plants <strong>or</strong> establish polym<strong>or</strong>phic populations.<br />

Theref<strong>or</strong>e, we expect that plant defenses that can provide signals f<strong>or</strong> herbiv<strong>or</strong>es could be superi<strong>or</strong> to defenses that have no<br />

,alue as signals. A game the<strong>or</strong>etical analysis is presented in <strong>or</strong>der to expl<strong>or</strong>e the evolution of plant defenses when plant<br />

]tness depends on the defensive status of the neighb<strong>or</strong>.<br />

LITERATURE CITED<br />

\TSATT, RR.and O'DOWD, D.J. 1976. Plant defense guilds. Science 193: 24-29.<br />

_.UGNER, M., FAGERSTROM, T., and TUOMI, J. 1991. Competition, defense and games between plants. Behav. Ecol.<br />

Sociobiol. 29:231-234.<br />

3OMZE, I.M. 1983. Lotka-Volterra equation and replicat<strong>or</strong> dynamics: A two-dimensional classification. Biol. Cybernetics<br />

48:201-211.<br />

)ANELL, K., EDENIUS, L., and LUNDBERG, R 1991. Herbiv<strong>or</strong>y and tree stand composition: Moose patch use in winter.<br />

Ecology 72: 1350-1357.<br />

ESHEL, I. 1972. On the neighb<strong>or</strong> effect and the evolution of altruistic traits. The<strong>or</strong>. Pop. Biol. 3: 258-277.<br />

3UILFORD, T. and CUTHILL, I. 1991. The evolution of aposematism in marine gastropods. Evolution 45:449-451.<br />

HAY, M.E. 1986. Associational plant defenses and the maintenance of species diversity: Turning competit<strong>or</strong>s into accomplices.<br />

Am. Nat. 128: 617-641.<br />

HJALTEN, J., DANELL, K., and LUNDBERG, R I993. Herbiv<strong>or</strong>e avoidance by association: Vole and hare utilization of<br />

woody plants. Oikos 68: 125-131.<br />

HOFBAUER, J. and SIGMUND, K. 1988. The The<strong>or</strong>y of Evolution and Dynamical Systems. Cambridge Univ. Press,<br />

Cambridge.<br />

LAUNCHBAUGH, K.L. and PROVENZA, F.D. 1993. Can plants practice mimicry to avoid grazing by mammalian<br />

herbiv<strong>or</strong>es? Oikos 66: 501-504.<br />

MATTSON, W.J., HERMS, D.A., WITTER, J.A., and ALLEN, D.C. 1991. Woody plant grazing systems: N<strong>or</strong>th American<br />

outbreak foliv<strong>or</strong>es and their host plants, p. 53-84. In Baranchikov, Y., Mattson, W.J., Hain, ER, and Payne, T.L., eds.<br />

F<strong>or</strong>est insect guilds: patterns of interaction with host plants. Gen. Tech. Rep. NE-153. Radn<strong>or</strong> PA: U.S. Department<br />

of Agriculture, F<strong>or</strong>est Service. 400 p.<br />

McNAUGHTON, S.J. 1978. Serengeti ungulates: Feeding selectivity influences the effectiveness of plant defense guilds.<br />

Science 199: 806-807.<br />

NEE, S. 1989. Does Hamilton's rule describe the evolution of reciprocal altruism? J. The<strong>or</strong>. Biol. 141: 81-91. "<br />

37


PFISTER, C.A. and HAY, M.E. 1988. Associational plant refuges: Convergent patterns in marine and terrestrial communities<br />

result from differing mechanisms. Oecologia 77: 118-129.<br />

PROVENZA, ED., PFISTER, J.A., and CHENEY, C.D. 1992. Mechanisms of learning in diet selection with reference to<br />

phytotoxicosis in herbiv<strong>or</strong>es. J. Range Manage. 45: 36-45.<br />

RHOADES, D.E 1979. Evolution of chemical defense against herbiv<strong>or</strong>es, pp. 3-54. In Rosenthal, G.A. and Janzen, D.H.,<br />

eds. Herbiv<strong>or</strong>es: Their Interactions with Secondary Plant Metabolites. Academic Press, New Y<strong>or</strong>k.<br />

ROSENBERG, G. 1991. Aposematism and synergistic selection in marine gastropods. Evolution 45:451-454.<br />

SABELIS, M.W. and de JONG, M.C.M. 1988. Should all plants recruit bodyguards? Conditions f<strong>or</strong> a polym<strong>or</strong>phic ESS of<br />

synomone production in plants. Oikos 53: 247-252.<br />

TAHVANAINEN, J.O. and ROOT, R.B. 1972. The influence of vegetational diversity of the population ecology of a<br />

specialized herbiv<strong>or</strong>e, Phyllotreta cruciferaea (Coleoptera: Chrysomelidae). Oecologia 10: 321-346.<br />

TAYLOR, RD. and JONKER, L.B. 1978. Evolutionary stable strategies and game dynamics. Math. Biosci. 40:145-1 56.<br />

TUOMI, J. and AUGNER, M. 1993. Synergistic selection of unpalatability in plants. Evolution 47: 668-672.<br />

WILSON, D.S. 1975. A the<strong>or</strong>y of group selection. Proc. Natl. Acad. Sci. USA 72: 143-146.<br />

ZEEMAN, E.C. 1981. Dynamics of the evolution of animal conflicts. J. The<strong>or</strong>. Biol. 89: 249-270.<br />

38


DEFENSE THEORIES AND BIRCH RESISTANCE<br />

MATTI ROUSI t, JORMA TAHVANAINEN 2, WILLIAM J. MATTSON 3<br />

and HANNI SIKANEW<br />

_The Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Punkaharju <strong>Research</strong> <strong>Station</strong>, SF-58450 Punkaharju, Finland<br />

2University of Joensuu, Depamnent of Biology, P.O. Box 111, SF 80101 Joensuu, Finland<br />

3<strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central Experimental <strong>Station</strong>, 1407 Harrison Rd, East Lansing, M148823 USA<br />

INTRODUCTION<br />

Different tree species in n<strong>or</strong>thern temperate f<strong>or</strong>ests clearly vary in their resistance to browsing mammals (Gill 1993<br />

a,b). In addition to variation at the species level, there are also clear indications of genetic differences in herbiv<strong>or</strong>e resistance<br />

within species. Such variation in resistance has been explained by the duration and intensity of past exposure to herbiv<strong>or</strong>es<br />

and pathogens (Leppik 1970, Bryant et al. 1989). It is believed that the centers of genetic diversity f<strong>or</strong> plant species should<br />

be the best places in which to find resistance to their common herbiv<strong>or</strong>es and diseases (Vavilov 1920). All of these kinds of<br />

variation have imp<strong>or</strong>tant economic and ecological implications f<strong>or</strong> plantation f<strong>or</strong>estry and f<strong>or</strong>est industries. F<strong>or</strong> example,<br />

hybridization could be used to elevate the resistance of otherwise high quality but susceptible tree species <strong>or</strong> individuals (see<br />

Rousi 1990).<br />

It has been postulated that there is a trade-off between growth and plant defense (Rhoades 1979, Herms and Mattson<br />

1992). That being the case, practical f<strong>or</strong>estry might not be interested in slow growing genotypes, although they might be<br />

m<strong>or</strong>e safe to use. Trees are generally attacked by a myriad of herbiv<strong>or</strong>es and if there is a trade-off between resistance and<br />

growth to each of those, then generally resistant genotypes should be very slow growing. Ecological trade-offs in the f<strong>or</strong>m<br />

of negative c<strong>or</strong>relations in resistance to different types of herbiv<strong>or</strong>es are also possible. Conventional wisdom suggests that<br />

trees which have to compete intensely f<strong>or</strong> nutrients and light, can probably be resistant only to restricted number of herbiv<strong>or</strong>es.<br />

In addition to genetic components, environment may also modify plant resistance. If carbon is the limiting fact<strong>or</strong> due<br />

to shading, <strong>or</strong> to large amounts of available nutrients and water, then defensive secondary metabolite level may decline so<br />

long as the plant still has strong sinks capable of usurping the available photosynthates and nutrients (Mattson 1980, Bryant<br />

el al. 1983).<br />

To test the effect of environment and birch genotype on hare resistance, we carried out several cafeteria experiments<br />

using winter d<strong>or</strong>mant shoots of small birch seedlings and plantlets grown in Punkaharju, East Finland.<br />

RESULTS AND DISCUSSION<br />

Birch Species and Genotypes Vary In Their Hare Resistance<br />

Birch species vary strongly in their hare resistance. Acc<strong>or</strong>ding to many studies, Japanese white birches, B.<br />

platyphylla and B. ermanii, are both very resistant. Some other species, such as B. alleghaniensis (Fig. 1) and B. papyrifera,<br />

are among the most susceptible species, whereas B. pendula is somewhat intermediate in resistance. Interestingly, experiments<br />

made in Alaska, Finland, and Japan all give essentially the same results (Chiba and Nagata 1968, Bryant et al. 1989,<br />

Rousi et al. 1989, 1996a). This most probably indicates the wide adaptability of birch: there are no drastic changes in<br />

resistance even if birches are grown as exotics. In addition, it shows that the Japanese mountain hare, Lepus ainu, the Finnish<br />

mountain hare, L. timidius, and the snowshoe hare, L. americanus, make their food selection on the same bases.<br />

Mattson, W.J., Niemel',i, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gem Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

39


P<br />

AL 70<br />

A 60<br />

T<br />

A 50<br />

B<br />

I<br />

I. 40<br />

I<br />

T 30<br />

Y<br />

20<br />

I<br />

N<br />

D 10<br />

E<br />

X 0<br />

B.alleghaniensis 16 x 21 15 x 21 17 x 21 B.pendula clone 39<br />

Figure I ._Hare palatabi]ity of B. alleghaniensis, three backcrossesof B. pendula andpla_phylla, local f<strong>or</strong>est <strong>or</strong>igin of B.<br />

pendul= andone micropropagatedc]oneof B. pendu{a. (::lone39 consistsof micropropagated plantlets of B. pendula.<br />

Mother trees (15,16 and 17) are hybrids between the two species and full sibs. Father trees (18,19 and 21) are South<br />

Finnish plus trees. Mean +S.E. Palatability index is the result from cafeteria experiments (two nights, six hares. F<strong>or</strong><br />

details see Rousi et al. 1991). Mean +S.E.<br />

Hybridization has been shown to be a very promising method to increase e.g., vole resistance of Larix species (Chiba<br />

et al. 1982). In addition, hybrids between susceptible and resistant birch species have been intermediate in their resistance to<br />

hares (Chiba and Nagata 1969). In <strong>or</strong>der to make preliminary tests of the inheritance of resistance, we tested the resistance of<br />

a backcross between Japanese and European birch [a hybrid between a Japanese white birch (very resistant) and European<br />

white birch (intermediate) was crossed with several South Finnish birch genotypes].<br />

The Japanese white birch used as a parent tree in our crossings was always of the same genotype. Generally,<br />

resistance of hybrids was at the same level as was that of local f<strong>or</strong>est <strong>or</strong>igins of European white birch (Figs. 1 and 2). Hybrids<br />

between Japanese and Fennoscandian white birches have shown superi<strong>or</strong> juvenile growth as rep<strong>or</strong>ted by Johnsson<br />

(1966). But the first year growth of our backcrossed seedlings was the same as that of European white birch (see Figs. 3 and<br />

4). Obviously the genetic base of our crossings was too limited f<strong>or</strong> any far reaching conclusions.<br />

Birch species and <strong>or</strong>igins derived largely from Pleistocene refugia such as Alaska, Siberia <strong>or</strong> Japan, are supposed to<br />

be m<strong>or</strong>e resistant than ecologically similar birches derived from nonrefugia (Bryant et al. 1989, Rousi 1988). It is difficult to<br />

test <strong>this</strong> hypothesis because it is difficult to verify whether particular birch populations came from authentic "refugia" <strong>or</strong> not.<br />

M<strong>or</strong>eover, no one knows the kind of selection pressure such populations have faced since release from their refugia. The<br />

most resistant and the most susceptible species in our experiments are from Japan, a country that was minimally glaciated<br />

during the Pleistocene. M<strong>or</strong>eover, B. pendula <strong>or</strong>igins from Siberia (not glaciated) and Finland (heavily glaciated) are equally<br />

resistant to hares (Rousi et al. 1996a). Experiments with micropropagated plantlets have revealed very large variation in hare<br />

resistance among different B. pendula genotypes growing in South Finland. Clone 39 especially has proved to be highly<br />

resistant under various experimental conditions (Fig. 5, Rousi et al. 1996b).<br />

Fast Growing Birches Are Also Resistant<br />

We tested the resistance of persistence-adapted yellow birch, B. alleghaniensis, against fast growing, sh<strong>or</strong>t-lived<br />

white birches in <strong>or</strong>der to find out whether inherent growth strategy is related to hare resistance of the species. The first year<br />

(greenhouse) growth of yellow birch was less (ANOVA, Bonferroni t-test, p


P 45<br />

A<br />

L 40<br />

A<br />

T 35<br />

A<br />

B 30<br />

I<br />

L 25<br />

I<br />

T 20<br />

Y<br />

15<br />

i 10<br />

N<br />

D 5<br />

E<br />

X 0<br />

16x19 16x18 15x18 17x19 17x18 clone 39<br />

e 2.--Hare palatability of backcrosses between B. pendula and B. platyphylla. Mother trees (15,16 and 17)are hybrids<br />

between the two species and full sibs. Father trees (18,19 and 21) are South Finnish plus trees. F<strong>or</strong> m<strong>or</strong>e details see<br />

Figure 1. Mean +S.E.<br />

P<br />

A<br />

L 60<br />

A [] B.alleghaniensis<br />

T<br />

A<br />

50<br />

B<br />

I<br />

L<br />

40<br />

I 30 mB.pendula<br />

T [] hybrid [] hybrid<br />

Y<br />

20<br />

[] hybrid<br />

I r = -.794<br />

N 10 p = .109<br />

D<br />

E 0 _-- I I I I I I ; I : :<br />

X<br />

62 63 64 65 66 67 68 69 70 71 72 73<br />

SEEDLINGHEIGHT(cm)<br />

•e 3.--Relationship between seedling growth and hare palatability. Material is the same as in Figure 1- H is f<strong>or</strong> hybrid<br />

(backcross). Clone 39 is omitted because plantlets were ca. 2-yr-old; seedlings were 1-yr-old.<br />

41


p<br />

A 40<br />

L<br />

A m<br />

T u<br />

A 35<br />

B<br />

I<br />

k<br />

i 30<br />

T<br />

Y<br />

II<br />

t<br />

!1<br />

25<br />

i r = -.580<br />

N<br />

D<br />

p = .305<br />

E 20 ..................... t -- : : : : , ,_ ; _ "<br />

X 63 64 65 66 67 68 69 70 71 72 73<br />

SEEDLING HEIGHT (cm)<br />

Figure 4.--Re|ationship between seedling growth and hare resistance. Material the same as in Figure 2. Clone 39 is<br />

omitted because plantlets were ca. 2-yr-old; seedlings were l-yr-old.<br />

P<br />

A 60<br />

L<br />

A<br />

T 50<br />

A<br />

B<br />

40<br />

I<br />

L<br />

! 30<br />

T<br />

Y<br />

20<br />

!<br />

N 10<br />

D<br />

E<br />

X 0<br />

K 1889 109 39<br />

CLONES<br />

Figure 5.--Hare resistance of three micropropagated genotypes of B. pendula. Clone 109 stands f<strong>or</strong> f. bircalensis (lobed<br />

leaves). Cafeteria experiments as in Figurel. Means +S.E.<br />

42


However, hares showed strong preference f<strong>or</strong> it. The general trend in our experiments has been that species and genotypes<br />

resistant to hares showed better than average growth, although the relationship was not statistically significant (Figs. 3 and 4).<br />

These results supp<strong>or</strong>t our earlier findings which indicate either a positive relationship between first year growth of birch<br />

seedlings and their hare resistance, <strong>or</strong> no relationship at all (Rousi et al. 1989, 1990, 1991, 1996b).<br />

European white birch genotypes have not shown any relationship between their growth and resistance to hares,<br />

Microtus voles, Curculionidae weevils, <strong>or</strong> Melamps<strong>or</strong>idium leaf rust (Poteri and Rousi 1996, Rousi et al. 1996b). These<br />

experiments did not supp<strong>or</strong>t any ecological trade-off either: resistance to any one pest was not related to resistance to others.<br />

However, vole and hare resistance in some experiments have shown a positive relationship (Chiba and Nagata 1969), and<br />

have also tended to do so in our own experiments (Rousi et al. t989, 1996b). Interestingly, clone 39 has shown generally<br />

high resistance to both biotic and abiotic threats (Fig. 5, Rousi et al. 1996b). The same clone also seems to be exceptionally<br />

resistant to ozone injury (clone 5-M, in P_ifikk6nen et al. 1993), whereas the generally herbiv<strong>or</strong>e susceptible clone 36 is<br />

similarly very susceptible to ozone damage (2-M, in P_i_ikk6nenet al. 1993).<br />

Environment May Sometimes Affect Birch Resistance<br />

Our previous experiments with various birch species (8 birch species) and <strong>or</strong>igins have indicated, as expected, that<br />

growth environment (shading <strong>or</strong> fertilization) does not have an effect on low resistance species (slow <strong>or</strong> fast growing), such<br />

as B. papyrifera, pubescens, maximowicziana, schmidtii <strong>or</strong> grossa. In high resistance species, fertilization increased the<br />

palatability of two species, B. platyphylla and B. ermanii, but did not affect another, B. resinifera. Shading had no clear<br />

effect on any of the tested species (Rousi et al. 1996a).<br />

European and Japanese white birches are closely related (Dugle 1966). However, fertilization did not diminish the<br />

hare resistance of European white birch (Rousi et al. 1991, 1993), but lowered that of Japanese white birch (Rousi et al.<br />

1996a). On the other hand, shading tends to increase hare palatability of European white birch (Figs. 6 and 7), but its effect<br />

on Japanese white birch is negligible (Rousi et al. 1996a). The effect of shading on European white birch is, however,<br />

strongly dependent on genotype and site fertility. There were no statistical differences in resistance among three families that<br />

we tested in low fertility soils (Fig. 6). But, if seedlings are grown in high fertility soil, resistance was significantly different<br />

in two out of three families (ANOVA, Bonferroni t-test, p


P FERTILIZATION E<br />

A<br />

L (<br />

A 0 - no shade<br />

T 60 50 = 50% shade (<br />

A<br />

B 50<br />

I 40<br />

L<br />

I 30<br />

T<br />

y 20<br />

10<br />

I<br />

N 0<br />

D<br />

E<br />

0 50 0 50 0 50<br />

X V 5845 x V 5848 V 5845 x V 5781 V 5845 x V 5846<br />

Figure 7.--Hare resistance of three B. pendula F2 families in two shade treatments. Seedlings were grown in nursery peat<br />

and obtained optimum fi3rtilization (F<strong>or</strong> details see Rousi et al. 1996a). Cafeteria experiments as in Figure 1. Mean +<br />

S.E.<br />

SUMMARY<br />

We conclude that prevailing defense ttle<strong>or</strong>ies, (e.g., Carbon-Nutrient Balance, Growth-Differentiation Balance, etc.)<br />

may not be fully applicable to birch. The reaction of birch species, <strong>or</strong>igins, andgenotypes to variable environments produces<br />

variable resurts, which depend on plant genotype, environment:and the particular mechanism of resistance to each pest type.<br />

Consequently, generalized conclusions about birch resistance should be made with great caution.<br />

LITERATURE CITED<br />

BRYANT, J.R, CHAPIN, F.S. [[I and KLEIN, D.R. 1983. Carbon/nutrient balance of b<strong>or</strong>eal plants in relation to vertebrate<br />

herbiv<strong>or</strong>y. Oikos 40:357-368.<br />

BRYANT, J.R, TAHVANAINEN, J., SULKINOJA, M., JULKUNEN-THTTO, R., REICHARDT R, and GREEN, T. 1989.<br />

Biogeographic evidence f<strong>or</strong> the evolution of chemical defense against mammalian browsing by b<strong>or</strong>eal birch and<br />

willow. The American Naturalist 134: 20-34.<br />

CHIBA, S. and NAGATA, Y. 1968. Resistance to field mice and mountain hares in Betula-species and hybrids. (1) Comparison<br />

between species. Qji Institute t\)r F<strong>or</strong>est Tree Improvement. Technical Note 68:20-24 (In Japanese).<br />

CttIBA, S. and NAGATA, Y. 1969. Resistance to field mice and mountain hares in Betula-species and hybrids. (2) Comparison<br />

between parental species and hybrids. Oji Institute f<strong>or</strong> F<strong>or</strong>est Tree Improvement. Technical Note 88: 153-<br />

155 (In Japanese).<br />

CHIBA, S., NAGATA, Y., and TOMAKL T 11982. Variation of vole resistance in Larix leptolepis and comparison with the<br />

hybrid L gmelini and L leptolepis. Oji Institute f<strong>or</strong> F<strong>or</strong>est Tree Improvement. Technical Note 209:6-9 (In Japanese).<br />

44


DUGLE, J.A. 1966. A taxonomic study of Western Canadian species in the genus Betula. Canadian Journal of Botany 44:<br />

929-1007.<br />

GILL, R.M.A. 1992a. A review of damage by mammals in n<strong>or</strong>th temperate f<strong>or</strong>ests: 1. Deer. F<strong>or</strong>estry 65(2): 145-169.<br />

GILL, R.M.A. 1992b. A review of damage by mammals in n<strong>or</strong>th temperate f<strong>or</strong>ests: 2. Small mammals. F<strong>or</strong>estry 65(3):<br />

281-308.<br />

HERMS, D.A. and MATTSON, W.J. 1992. The dilemma of plants: to grow <strong>or</strong> defend. The Quarterly Review of Biology<br />

67(3): 283-335.<br />

JOHNSSON, H. 1966. Avkommepr6vning av bj6rk. F6reningen Skokstrfidsf6r_idling, Arsbok 1966: 90-135.<br />

LEPPIK, E.E. 1970. Gene centers of plants as a source of disease resistance. Annual Review of Phytopathology 8: 323-344.<br />

MATTSON, W.J. 1980. Herbiv<strong>or</strong>y in relation to plant nitrogen content. Annual Review of Ecology and Systematics 11:<br />

119-161.<br />

PAAKKONEN, E., PAASiSALO, S., ttOLOPAINEN, T., and K._RENLAMPI, L. Growth and stomatal responses of birch<br />

(Betula pendula Roth.) clones to ozone in open-air and chamber fumigations. New Phytologist 125: 615-623.<br />

POTERI, M. and ROUSI, M. 1996. Variation in Melamps<strong>or</strong>idium resistance among European white birch clones grown in<br />

different fertilizer treatments. Eurp. J. F<strong>or</strong>. Pathol.: In Press.<br />

RHOADES, D.F. 1979. Evolution of plant chemical defense against herbiv<strong>or</strong>es, p. 3-54. In Rosenthal, G.A. and Janzen<br />

D.H., eds. Herbiv<strong>or</strong>es: their interaction with secondary plant metabolites. Academic Press, Orlando.<br />

ROUSI, M. 11988. Resistance breeding against voles in birch: possibilities f<strong>or</strong> increasing resistance by provenance transfers.<br />

EPPO (European Plant Protection Organization) Bulletin 18: 257-263.<br />

ROUSI, M. 1990. Breeding f<strong>or</strong>est trees f<strong>or</strong> resistance to mammalian herbiv<strong>or</strong>es - a study based on European white birch.<br />

Acta F<strong>or</strong>estalia Fennica 210.<br />

ROUSI, M., TAHVANAINEN, J., and UOTILA, I. 1989. Inter- and intraspecific variation in the resistance of winterd<strong>or</strong>mant<br />

birch (Betula spp.) against browsing by the mountain hare. Holarctic Ecology 12:187-192.<br />

ROUSI, M., HENTTONEN, H., and KAIKUSALO, A. 1990. Resistance of birch (Betula pendula and B. platyphylla)<br />

seedlots to vole. Scandinavian Journal of F<strong>or</strong>est <strong>Research</strong> 5: 427-436.<br />

ROUSI, M., TAHVANAINEN, J., and UOTILA, I. 1991. A mechanism of resistance to hare browsing in winter-d<strong>or</strong>mant<br />

European white birch (Betula pendula). The American Naturalist 137: 64-82.<br />

ROUSI, M., TAHVANAINEN, J., and UOTILA, I. 1993. Effects of shading and fertilization on resistance of winterd<strong>or</strong>mant<br />

birch (Betula pendula) to voles and hares. Ecology 74(1): 30-38.<br />

ROUSI, M., MATTSON, W.J., TAHVANAINEN, J., HENTTONEN, H., KOIKE, T., and UOTILA, I. 1996a. Growth and<br />

hare resistance of birches: testing defense the<strong>or</strong>ies. Oikos: In Press.<br />

ROUSI, M., TAHVANAINEN, J., HERMS, D.H, HENTTONEN, H., and UOTILA, I. 1996b. Clonal variation in susceptibility<br />

of white birches to herbiv<strong>or</strong>es. F<strong>or</strong>. Sci.: Submitted.<br />

VAVILOV, N.I. 1922. The law of homologous series in variation. Journal of Genetics 12(1): 47-89.<br />

45


A REDOX-BASED MECHANISM BY WHICH<br />

ENVIRONMENTAL STRESSES ELICIT CHANGES IN<br />

PLANT DEFENSIVE CHEMISTRY<br />

DALE M. NORRIS<br />

642 Russell Lab<strong>or</strong>at<strong>or</strong>ies, University of Wisconsin, Madison, WI 53706, USA<br />

INTRODUCTION<br />

It is now clear that most, if not all, lifef<strong>or</strong>ms employ chemicals in communicating with their biotic and abiotic<br />

environments. Higher plants and animals, including insects and humans, may even use many of the same compounds<br />

(Rodriquez and Levin 1976, N<strong>or</strong>ris and Liu 1992, Raina et aL 1992). Quinol-quinone redox couples, e.g., p-hydroquinone/<br />

p-benzoquinone (Fig. 1), are prime examples of such common messengers. Hundreds of ions, free radicals (Packer and<br />

Glazer 1990) and molecules have been shown to be messengers. Alkaloids, flavonoids and terpenoids are families of<br />

compounds which are especially imp<strong>or</strong>tant as messengers (N<strong>or</strong>ris 1986).<br />

0 O- OH<br />

II<br />

II II<br />

0 0 O- OH<br />

"A" "B"<br />

Figure l.--The classical redox couple in physical chemistry, p-hydroquinone / p-benzoquinone.<br />

The development of knowledge about the chemical senses has frequently been quite disconnected among many<br />

scientific disciplines. As a result, limited progress has been made in finding unifying principles. However, f<strong>or</strong> several<br />

decades, I have developed hypotheses pertaining to such unifying principles and tested them experimentally (N<strong>or</strong>ris 1994).<br />

The results of our and others' eff<strong>or</strong>ts have allowed me to propose that a common code f<strong>or</strong> chemical communication exists<br />

among lifef<strong>or</strong>ms, and them and their abiotic environments. The proposed name f<strong>or</strong> <strong>this</strong> code is the Environmental Energy<br />

Exchange (EEE) Code (N<strong>or</strong>ris 1981, 1986, 1988, 1994; Neupane and N<strong>or</strong>ris 1992; N<strong>or</strong>ris and Liu 1992).<br />

In <strong>this</strong> paper, the essential characteristics of <strong>this</strong> EEE code, including its underlying sulfhydryl (thiol, -SH) / disulfid_<br />

(-S-S-) -dependent oxidative-reductive, energy-transducing mechanism, are applied to the interpretation of how environmen-<br />

tal stresses elicit changes in plant defensive chemistry.<br />

Stress Alteration Of Phytoehemieal Messages<br />

The chemistry by which plants "communicate" with their herbiv<strong>or</strong>es has been classified as either constitutive<br />

(i.e., pre-stress) <strong>or</strong> inducible (i.e., post-stress). Plant pathologists and plant physiologists created <strong>this</strong> classification to<br />

Mattson, W.J., Niemel_i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. USD_-X<br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC- 183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

46<br />

!


accommodate findings related to microbial interactions with plants (Bell 1981). Constitutive defensive phytochemistry<br />

is that expressed in the plant by its genome in the supposed absence of specific stress from the environment; whereas,<br />

inducible defensive phytochemistry is that expressed in response to a specific stress caused by some fact<strong>or</strong> (e.g., microbial<br />

infection) in the plant's environment. Although these interpretations were developed f<strong>or</strong> plant-microbe interactions,<br />

subsequent studies, especially by entomologists and chemical ecologists, have shown that much of the involved phytochemistry,<br />

and some of the developed concepts, may prove useful to understanding messenger phytochemicals in<br />

herbiv<strong>or</strong>e-plant communications (Kogan and Paxton 1983). Our studies of such phytochemistry indicate that the maj<strong>or</strong><br />

differences between pre- and post-stress chemistry are largely quantitative (e.g., Neupane and N<strong>or</strong>ris 1991a, Markovic et<br />

al. 1993). Thus, a plant usually responds to environmental stresses by making m<strong>or</strong>e, <strong>or</strong> less, of individual compounds in<br />

a given (genome-determined) array of secondary metabolites. These chemicals have sometimes been termed 'phytoalexins'<br />

by plant pathologists and physiologists.<br />

Unifying Mechanisms Of Elieitation<br />

A wide diversity of biotic and abiotic entities have proven effective as exo-elicit<strong>or</strong>s (external to the plant cell) in the<br />

inducible alteration of plant defensive chemistry against environmental stresses (Sequeira 1983). However, plant pathologists<br />

and physiologists have especially focused on fragments of microbial-pathogen <strong>or</strong> plant-cell walls as exo-elicit<strong>or</strong>s. _3-glucans<br />

from ptant-cel] walls apparently play maj<strong>or</strong> elicit<strong>or</strong>y roles (Ryan 1983, Darvill and Albersheim 1984, Templeton and Lamb<br />

1988, Dixon and Lamb 1990). Heavy metals, including mercury, and sulthydryl reagents are other proven exo-elicit<strong>or</strong>s<br />

(Stossel 1984). Although the diversity of exo-elicit<strong>or</strong>s might imply that a common mechanism of exo-elicitation is unlikely,<br />

the suggested involvements of sulfhydryls in the process led Sequeira (1983) to hypothesize their imp<strong>or</strong>tance in such inductions.<br />

Other rep<strong>or</strong>ts of elicitation agents reacting with sulfhydryls in phytoalexin biosynthesis (Stossel 1984) supp<strong>or</strong>t the<br />

underlying imp<strong>or</strong>tance of sulfhydryls in elicitation processes. M<strong>or</strong>e recent studies (Neupane and N<strong>or</strong>ris (1990, 1991a) and<br />

Liu et al. (1992, 1993)) showed that several classical compounds f<strong>or</strong> detecting (i.e., reacting with) sulfhydryls also alter the<br />

biosynthesis of isoflavanoid phytoalexins in soybeans, Glycine max (L.) Merrill. These induced phytoalexins also alter the<br />

defense of these plants to some insect herbiv<strong>or</strong>es (Sharma and N<strong>or</strong>ris 1991).<br />

It now seems clear that the elicit<strong>or</strong>y mechanism in plants involves directly, <strong>or</strong> indirectly, the oxidative-reductive<br />

(redox) chemistry of sulfhydryl (-SH) / disulfide (-S-S-) interactions (Fig. 2). First, reagents which react specifically with<br />

sulfhydryls are effective elicit<strong>or</strong>s of stress-alterable defensive phytochemistry (Stossel 1984; Sequeira 1983; Neupane and<br />

N<strong>or</strong>ris 1990, 1991a; Liu etal. 1992, 1993, 1994; Haanstad and N<strong>or</strong>ris 1992; N<strong>or</strong>ris and Liu 1992; N<strong>or</strong>ris 1994). Second,<br />

sulfhydryl proteins in the plasma membrane which surrounds each plant cell have been shown to react with elicit<strong>or</strong>y sulfhydryl<br />

reagents (e.g., p-chl<strong>or</strong>omercuriphenyl sulfonic acid, PMBS) when applied in vivo to the intact plant (Liu et al. 1992,<br />

1993, 1994). Such reaction, <strong>or</strong> interaction, between a sulfhydryl reagent and a sulfhydryl protein (i.e., recept<strong>or</strong>) at the plasma<br />

membrane level may trigger significant alterations in the defensive chemistry of elicited plant cells (Liu et al. 1992, 1994;<br />

N<strong>or</strong>ris 1994). Third, the essential mechanism by which sulfhydryl reagents and other proven elicit<strong>or</strong>y entities, both biotic<br />

and abiotic, function in altering defensive phytochemistry is apparently oxidative-reductive (Neupane and N<strong>or</strong>ris 1991b,<br />

1992; Markovic et al. 1993; N<strong>or</strong>ris and Liu 1992). The classical antioxidants, L-asc<strong>or</strong>bic acid (vitamin C) and o_-tocopherol<br />

(vitamin E), have proven to be effective elicit<strong>or</strong>s when applied to whole plants (Neupane and N<strong>or</strong>ris 1991b, 1992). Fourth,<br />

i Neupane and N<strong>or</strong>ris (1992) showed that L-asc<strong>or</strong>bic acid and herbiv<strong>or</strong>y by Trichoplusia ni (Hubner) elicited analogous<br />

defensive responses in soybeans. Thus, the involvement of a common mechanism of elicitation, i.e., oxidative-reductive, is<br />

supp<strong>or</strong>ted. Our current interpretation is that herbiv<strong>or</strong>y elicits inducible defensive phytochemistry through triggering oxidation<br />

/ reduction-dependent reactions which are common with those caused in plant tissues by a wide variety of other environ-<br />

! mental stresses.<br />

-$H /-5-5-<br />

Figure 2.--The sulfur-based redox couple, sulfhydryl (thiol, -SH) ! disulfide (-S-S-), in which the addition of reducing<br />

agent fav<strong>or</strong>s the f<strong>or</strong>mation of thiols; whereas, the addition of oxidizing agent fav<strong>or</strong>s the reverse, i.e., f<strong>or</strong>mation of<br />

disulfides. This redox couple in proteins is considered the functional interface by which energy is exchanged as<br />

inf<strong>or</strong>mation between living cells and their environments.<br />

47


METHODS<br />

Antioxidant Alteration Of Ash Antixenosis To Malacosoma disstria<br />

Because sulfhydryl groups may readily participate in redox reactions in living cells (M<strong>or</strong>ton 1965), our studies<br />

evaluated a natural antioxidant (reducing agent), o_-tocopherol(vitamin E), as an elicit<strong>or</strong> of defensive phytochemistry. Alphatocopherol<br />

was tested in a basal trunkband on green ash trees to alter their antixenosis to f<strong>or</strong>est tent caterpillar larvae,<br />

Malacosoma disstria (Hubner). The trees were 4.6m tall 'Summit' green ash, Fraxinus pennsylvanica var. subintegerrinia<br />

(Vahl) Fernald, and had a mean trunk diameter of 5.1 cm at 1 m above the soil surface. The basal trunk of each tree received<br />

o_-tocophero[, a proven elicit<strong>or</strong> (Neupaneand N<strong>or</strong>ris 1991b), as a band application at the dosages of 25.0 <strong>or</strong> 50.0 IU / ml in<br />

mineral oil. An international unit (IU) equals 1 mg of all-rac-cx-tocopheryl acetate (U.S. Pharmacopaea 1980). Sixty ml of<br />

either concentration of ot-tocopherol in mineral oil were placed on a 120cm 2bandage; the control trees received only 60 ml of<br />

mineral oil.<br />

Insect Bioassays<br />

Leaves f<strong>or</strong> bioassay were removed from three distinct trees per treatment <strong>or</strong> control at each of several intervals after<br />

treatment, and two-choice feeding assays were conducted with 1.5-cm-diam disks cut with a No. 8 c<strong>or</strong>k b<strong>or</strong>er from such<br />

leaves and with third-instar f<strong>or</strong>est tent caterpillars. The insect's feeding option thus was between a comparable leaf disk from<br />

an elicited versus a solvent-control tree. The quantitation of insect feeding was detailed by Markovic et al. (1993).<br />

Chemical Analyses<br />

[.,eaves f<strong>or</strong> chemical analyses were collected from three distinct ash trees f<strong>or</strong> each elicitation dosage and the solvent<br />

control, and f<strong>or</strong> 8 and 16days after treatment. Three compound leaves from each tree were immediately put individually into<br />

a glass jar containing 80% methanol, and then st<strong>or</strong>ed in darkness at -20°C until chemical analysis. Procedures used to extract<br />

chemicals from ash leaves; hydrolyze extracted chemicals; and analyze the chemicals by high perf<strong>or</strong>mance thin layer<br />

chromatography (HPTLC) and high perf<strong>or</strong>mance liquid chromatography (HPLC) were as detailed by Markovic et al. (1993).<br />

RESULTS<br />

Antioxidant Altered Foliar Antixenosis In Ash<br />

F<strong>or</strong>est tent caterpillar herbiv<strong>or</strong>y was altered on leaf disks from green ash trees elicited with either of the two dosages<br />

of c_-tocopherol as compared to disks from solvent-treated ash trees (Table 1). High perf<strong>or</strong>mance thin layer chromatography<br />

(HPTLC) revealed distinct differences in both the non-hydrolyzed and hydrolyzed p<strong>or</strong>tions of the ethyl acetate extractables<br />

from elicited versus control ash trees (Markovic et al. 1993). Reduced insect preference f<strong>or</strong> foliage due to tocopherol<br />

elicitation was accompanied by an increased mean total HPLC-resolved peak area of ethyl acetate extractables from the nonhydrolyzed<br />

leaf sample as compared to control foliage. Whereas, increased insect preference f<strong>or</strong> foliage due to elicitation<br />

was accompanied by a decreased total HPLC-resolved peak area of such extractables as compared to control. HPLC of the<br />

above non-hydrolyzed, ethyl acetate extractables showed mainly quantitative, rather than qualitative, differences between ¢xtocopherol-elicited<br />

versus solvent-treated controls. The differences were especially evident in five maj<strong>or</strong>, and three lesser,<br />

peaks; these eight peaks thus were chosen f<strong>or</strong> a m<strong>or</strong>e detailed comparison between the non-hydrolyzed fraction of the ethyl<br />

acetate extractables from the two cx-tocopherol treatments and solvent-treated controls at two times (8 and 16 days) after<br />

elicitation.<br />

At 8 days after elicitation, the foliage from trees receiving 25 IU / ml contained significantly m<strong>or</strong>e of HPLC peaks<br />

1-6, but not of peaks 7 <strong>or</strong> 8, than did leaves from solvent-control trees (Fig. 3). Foliage from these treated trees also was<br />

significantly less preferred than that from the control trees (Table 1). Conversely, at 16 days after elicitation, foliage from<br />

trees that received either 25 <strong>or</strong> 50 IU / ml was preferred over that from the solvent-control ones (Table 1). The ethyl acetate<br />

extractables at 16 days after elicitation from preferred foliage from trees that received 25 <strong>or</strong> 50 IU / ml had a smaller average<br />

total HPLC-resolved peak area than did those from the leaves of solvent-control trees (Fig. 4). The mean total HPLC-<br />

resolved peak area was also significantly different between the foliage collected from the solvent-control trees at 8 and 16<br />

days after elicitation. Thus, time in thegrowing season also affected the chemical composition of the trees.<br />

48


Table l.--Mean ±S.E. area eaten by the f<strong>or</strong>est tent caterpillar in leaf disks from tocopherol-elicited versus<br />

control-elicited ash trees.<br />

Behavi<strong>or</strong>al Area eaten (cm2) b<br />

response Days _ Treatment Control<br />

Nonpreferred _ 8 0.18 _+0.04 "*c 0.41 + 0.04<br />

Preferred 1_ 16 0.49 + 0.07* 0.31 + 0.06<br />

Preferred 2d t6 0.46 + 0.05"" 0.16 + 0.04<br />

Days after elicitation.<br />

b From Markovic et al. (1993).<br />

Vitamin E dose was 25.0 IU / ml.<br />

d Vitamin E dose was 50.0 tU / ml.<br />

Means followed by * <strong>or</strong> **are significantly different from their control at p < 0.05 and p < 0.01, respectively.<br />

35000<br />

30000<br />

25000<br />

R Non-preferred<br />

a E! Control B<br />

b a<br />

W<br />

¢ 20000 b<br />

w 15000<br />

a b<br />

10000 b a<br />

a<br />

a<br />

a a<br />

5000 b a a<br />

0<br />

1 2 3 4 5 6 7 8<br />

PEAK NUMBER<br />

gure 3.--Comparative mean areas of the indicated peaks in the standardized HPLC analysis of the nonhydrolyzed fraction<br />

of the ethyl acetate extractables from ash foliage. Data are f<strong>or</strong> nonpreferred foliage (elicited with 25 IU / ml) and<br />

control B foliage (elicited only with the solvent). Leaves were collected at 8 days after elicitation. Peak areas not<br />

followed by the same letter are significantly different at p < 0.05.<br />

49


35000 j<br />

30000<br />

25000 b b<br />

<<br />

w 20000<br />

IZ: <<br />

b<br />

m 15000 b<br />

O.<br />

10000<br />

b<br />

m Preferred 1<br />

a IW Control A<br />

r'l Preferred 2<br />

a a a b<br />

5000 c a<br />

0 . . .<br />

1 2 3 4 5 6 7 8<br />

PEAK NUMBER<br />

Figure 4.--Comparative mean areas of the indicated peaks in the HPLC analysis of the nonhydrolyzed fraction of ethyl<br />

acetate extractables from ash foliage. Data are f<strong>or</strong> preferred 1 (elicited with 25 IU / ml), preferred 2 (elicited with 50<br />

IU / ml and control A (elicited only with solvent) foliage. Leaves were removed from these trees at 16 days after<br />

elicitation. Peak areas not followed by the same letter are significantly different at p < 0.05.<br />

DISCUSSION<br />

Antioxidant Altered Foliar Antixenosis<br />

The findings supp<strong>or</strong>t the interpretationthat preference / non-preference decisions by M. disstria larvae regarding __<br />

tocopherol-elicited versus solvent (control)-elicited green ash foliage are probably attributable to quantitative changes in<br />

several compounds, not just in one chemical. Elicitation by _-tocopherol alters the quantitative contributions of at least<br />

several chemicals in the mixture of secondary metabolites. The effects are very dynamic, e.g., changing from increased to<br />

decreased antixenosis between 8 and 16 days after elicitation with 25 IU / ml. These findings seem compatible with the<br />

knowledge that surviv<strong>or</strong> plants systematically switch their maj<strong>or</strong> allocation of nutrient and energy resources between primary<br />

growth and secondary growth-differentiation including defensive chemistry. Such dynamic switching seems reasonable<br />

because surviv<strong>or</strong> plants must grow, but they must also defend themselves from potentially lethal environmental stresses,<br />

50<br />

b<br />

i


Sulfhydryl-Disulfide Mechanisms In Plant And Animal Perceptions<br />

Most, if not all, common chemical messengers between insects and plants, and between them and their environments<br />

(Rodriguez and Levin 1976, Neupane and N<strong>or</strong>ris 1992, N<strong>or</strong>ris and Liu 1992, Raina et al. 1992) are both elicited and perceived<br />

by sulfhydryl / disulfide (-SH / -S-S-) -dependent mechanisms (Neupane and N<strong>or</strong>ris 1992, Liu et al. 1992, N<strong>or</strong>ris<br />

1994). The energy-transduction mechanism in chemical communications by both animals and plants is sulfur dependent.<br />

The initial studies in <strong>this</strong> area showed that 1,4-naphthoquinones serve as repellents and deterrents to Periplaneta americana<br />

L., the American cockroach, and Scolytus multistriatus (Marsh.), the smaller European elm bark beetle, by reacting with<br />

sulfhydryls in recept<strong>or</strong> proteins in dendritic membranes of chemosensitive neurons (N<strong>or</strong>ris et al. 1970a, 1970b, 1971;<br />

Rozental and N<strong>or</strong>ris 1973, 1975; Singer et al. 1975).<br />

The redox dependency of chem<strong>or</strong>eception in insects was also shown in studies (N<strong>or</strong>ris 1969, 1970) where p-hydroquinone,<br />

the reduced partner in the classical redox couple, p-hydroquinone / p-benzoquinone, excited feeding by S.<br />

multistriatus; whereas, the oxidized partner, p-benzoquinone, inhibited feeding. In spite of these early findings, enthusiasm<br />

f<strong>or</strong> redox-based recept<strong>or</strong> and energy-transduction mechanisms in animal and human chem<strong>or</strong>eception has, until quite recently,<br />

been limited. This was especially true f<strong>or</strong> insect chem<strong>or</strong>eception. Although numerous research rep<strong>or</strong>ts have confirmed that<br />

insect chem<strong>or</strong>eception is sulfhydryl-disulfide dependent (Villet 1974; Frazier and Heitz 1975; Singer et al. 1975; Ma 1977,<br />

1981; Vande Berg 1981), several w<strong>or</strong>kers in <strong>this</strong> entomological specialty (e.g., Kaissling 1971, 1974, 1987; Vogt and<br />

Riddifbrd 1986) concluded that non-covalent bonding, but not redox reactions, are involved in the transduction mechanism in<br />

insect chern<strong>or</strong>eception. Until recently, most researchers w<strong>or</strong>king in recept<strong>or</strong> and energy-transduction mechanisms in living<br />

cells also had concluded that redox reactions were not involved. In contrast, findings by N<strong>or</strong>ris (1969, 1971, 1976, 1981,<br />

1985, 1988, 1994) have consistently indicated that redox chemistry is fundamentally involved in recept<strong>or</strong> function and<br />

energy-transduction mechanisms. The validity of our results has been reconfirmed throughout the general field of recept<strong>or</strong>s<br />

and signal transduction (e.g., St<strong>or</strong>z et al. 1990, Van Der Vliet and Bast 1992, Stamler et al. 1992, Reichard 1993, Pyle 1993,<br />

Ravichhandran et al. 1993). The sulfhydryl-disulfide dependent:redox chemistry of recept<strong>or</strong>s and energy transduction has<br />

become a maj<strong>or</strong> area of research throughout biology (Van Der Vliet and Bast 1992). Thus, these aspects of insect chem<strong>or</strong>eception<br />

have not only been confirmed, but also are now widely viewed as a part of a much larger chem<strong>or</strong>eception 'whole'<br />

within biology.<br />

The Required Elemental Trait<br />

The fundamental trait required f<strong>or</strong> the evolution of a system f<strong>or</strong> inf<strong>or</strong>mation generation, transfer and use in maintaining<br />

<strong>or</strong>der is the transfer of chemical groups and energy. This required trait f<strong>or</strong> chemical messengers singles out phosph<strong>or</strong>us<br />

and sulfur from all other atoms in the "Periodic System" (Wald 1969). The two basic atomic characteristics which qualify<br />

sulfur and phosph<strong>or</strong>us uniquely as agents of chemical group and energy transfers are (a) the possession of d, in addition to s<br />

and p, <strong>or</strong>bitals which increases their capacities to f<strong>or</strong>m linkages with a variety of energy potentials and (b) an intrinsic<br />

instability of such linkages, which facilitates the exchange of chemical groups and energy (i.e., inf<strong>or</strong>mational units) (Wald<br />

1969). Such roles f<strong>or</strong> phosph<strong>or</strong>us are well established experimentally, but they are only just now being recognized f<strong>or</strong> sulfur.<br />

However, sulfur now seems to be the most highly qualified element regarding abilities to exchange energy as inf<strong>or</strong>mation<br />

among living f<strong>or</strong>ms and their abiotic environments (Wald 1969; N<strong>or</strong>ris et al. 1970a, 1970b, 1971; N<strong>or</strong>ris 1971, 1979,<br />

1981, 1985, 1986, 1988, 1994; Neupane and N<strong>or</strong>ris 1992; N<strong>or</strong>ris and Liu 1992; Van Der Vliet and Bast 1992). First of all,<br />

regarding such roles, sulfur, as thiol (-SH), brings to living systems a f<strong>or</strong>m of <strong>or</strong>ganic sulfur which possesses those atomic<br />

traits which are otherwise limited to <strong>or</strong>ganic oxygen and are essential to <strong>or</strong>ganisms. However, thiol lacks most, if not all, of<br />

those traits (<strong>or</strong> degrees thereof) possessed by oxygen which make oxygen <strong>or</strong> its free-radical derivatives, especially harmful,<br />

even deadly to living <strong>or</strong>ganisms, unless the oxygen is handled in special ways, e.g., bound to hemoglobin in vertebrate blood<br />

(Wald 1969, Harold 1986). Second, sulfur as well as phosph<strong>or</strong>us f<strong>or</strong>ms relatively long (i.e., loose) bonds with other atoms;<br />

such bonds hold the attached atom less tightly and thus it may be m<strong>or</strong>e readily transferred, exchanged, as inf<strong>or</strong>mational<br />

energy in living systems (Wald 1969, Harold t986). Third, the trait which ultimately sets sulfur, as a vehicle of chemical<br />

group and energy transfers as inf<strong>or</strong>mation, apart from phosph<strong>or</strong>us is the f<strong>or</strong>mer's ability to f<strong>or</strong>m a thiol (-SH). This property<br />

uniquely enables sulfur to readily accept, donate <strong>or</strong> variously share the basic energy unit in living systems, hydrogen (Wald<br />

1969, Szent-Gy<strong>or</strong>gyi 1973). Thus, through sulfhydryl / disulfide (thiol, -SH / disulfide, -S-S-) redox chemistry, sulfur<br />

uniquely brings to living cells, and their chemical communications, a "yin and yang", a give and take, mechanism f<strong>or</strong> moving<br />

hydrogens, the basic energy unit of living systems, as inf<strong>or</strong>mation between plants and animals and their abiotic environments<br />

(N<strong>or</strong>ris 1986, 1988, 1994; Neupane and N<strong>or</strong>ris 1992; N<strong>or</strong>ris and Liu 1992).<br />

51


The conversion of abiotic inf<strong>or</strong>mational energy states, as in pheromone and phytochemical messengers, into biotic<br />

inf<strong>or</strong>mational energy states occurs in sulfhydryl / disulfide-proteinaceous recept<strong>or</strong>s in the plasma membrane which encloses<br />

each living cell (N<strong>or</strong>ris 1981, 1986, 1988). Chemical messengers, by directly <strong>or</strong> indirectly oxidizing sulfhydryls <strong>or</strong> reducing<br />

disulfides in proteins of living cells, allow the folding and unfolding, respectively, of the three-dimensional structure of such<br />

macromolecules in cells (Figs. 5) (Sela et al. 1957, N<strong>or</strong>ris 1981). Such folding and unfolding of proteins (e.g., as in muscles<br />

during their contractions versus relaxations, <strong>or</strong> in nerve membranes during impulse generation versus decay) especially bring<br />

to living systems, motion; a trait which is so commonly associated with life (Harold 1986).<br />

tt -<br />

(<br />

H,,,- } _<br />

Figure 5.uA schematic illustrating how the f<strong>or</strong>mation of disulfide (-S-S-) bonds in proteins can stabilize the three-dimensional<br />

structure of such macromolecules in a m<strong>or</strong>e aggregated state (see on the right); whereas, the reduction of those<br />

-S-S- bonds by reducing agents (e.g., some chemical messengers) can yield sulfhydryls (-SHs) and a m<strong>or</strong>e "opened",<br />

non-aggregated, three-dimensional f<strong>or</strong>m to the protein in the recept<strong>or</strong> and energy-transducer macromolecule (see on<br />

the/eft). In <strong>this</strong> illustration, each circle in the chains of circles represents an amino acid in the chain of amino acids<br />

which make up a protein. Each circle containing a sulfur (S) represents the amino acid, cysteine, which bears one<br />

sul thydryl (-SH).<br />

As N<strong>or</strong>ris and cow<strong>or</strong>kers (N<strong>or</strong>ris et al. 1970a, 1970b, 1971; Rozental and N<strong>or</strong>ris 1973; Singer et al. 1975) showed<br />

about 20 years ago, the highly reversible redox system based on sulfhydryl / disulfide interactions with chemical messengers<br />

yields not only qualitative (i.e., excitat<strong>or</strong>y <strong>or</strong> inhibit<strong>or</strong>y), but also quantitative (i.e., quantal, unitized) exchanges of energy.<br />

An inf<strong>or</strong>mational system requires simply an ability to map an output (i.e., efferent) message <strong>or</strong> behavi<strong>or</strong> on to an input (i.e ....<br />

afferent) message <strong>or</strong> stinmlus (O'Connell 1981). The above described highly reversible sulfhydryl / disulfide redox system<br />

f<strong>or</strong> transferring energy (hydrogens, <strong>or</strong> electrons and protons) as inf<strong>or</strong>mation is readily quantifiable in millivolts (mV)<br />

especially by polarography (Rozental and N<strong>or</strong>ris 1973, 1975) <strong>or</strong> the electroantennogram (EAG) (N<strong>or</strong>ris and Chu 1974; N<strong>or</strong>ris<br />

1979, 1986, 1988). Thus, <strong>this</strong> system fully satisfies the requirements of a code f<strong>or</strong> the inf<strong>or</strong>mational exchange of energy.<br />

The<strong>or</strong>ized Environmental Energy Exchange Code<br />

An inf<strong>or</strong>mational code thus must be based on (a) an universal medium (i.e., messenger) and (b) a means (method) of<br />

using that universal messenger to convey both qualitative and quantitative messages. Such a system allows a special use of<br />

energy f<strong>or</strong> establishing <strong>or</strong>der (Harold 1986). N<strong>or</strong>ris ( 1971, 1976, 1986, 1988, 1994) has presented the<strong>or</strong>ies, based on sulfur<br />

electrochemistry as the universal messenger, f<strong>or</strong> chemical communications among animals, including humans, and plants;<br />

and between them and their environments. The most recent the<strong>or</strong>etical version of such a code is here termed the 'Environ-<br />

mental Energy Exchange Code'. In the proposed EEE code, sulfur electrochemistry is the universal medium (messenger);<br />

and the thiot (-SH), its derivatives and its redox couple, disulfide (-S-S-), are the chemical means f<strong>or</strong> using sulfur electrochemistry<br />

to convey both qualitative and quantitative messages. Thus, through chemical-messenger oxidation of a sulfhydryl<br />

(thiol, -SH) in each of two molecules of the amino acid cysteine in the recept<strong>or</strong> protein (energy-transducer), a disulfide (-S-S-<br />

) may f<strong>or</strong>m (Fig. 2 ). Such a disulfide in the three-dimensional protein involves a conf<strong>or</strong>mational (shape) change; such<br />

energy transfer thus allows the protein to become m<strong>or</strong>e folded (aggregated) in shape (Fig. 2 ). Conversely, through chemicalmessenger<br />

reduction of a disulfide, the -S-S- bridge between the two cysteines is broken and the protein may then assume a<br />

m<strong>or</strong>e open (i.e., unfolded) three-dimensional shape (Fig. 2 ). This chemical messenger-driven, highly reversible sulfhydryl /<br />

disulfide electrochemistry in the recept<strong>or</strong> protein (energy-transducer) thus brings to <strong>or</strong>ganisms quantitative dynamic f<strong>or</strong>m,<br />

which is so essential to function.<br />

52


Iraa pri<strong>or</strong> experimental analysis of such inf<strong>or</strong>mation exchange between the feeding-inhibit<strong>or</strong>y messenger menadione<br />

(2-methyt-t ,4-naphthoquinone) from the environment, and sulthydryls in the dendritic membrane of chemosensitive neurons<br />

in the antenna of Periplaneta americana, the electrochemical transduction of inf<strong>or</strong>mational energy into <strong>or</strong>ganismal response<br />

(i.e., activity) was described mathematically as a regression with r= 0.997 (N<strong>or</strong>ris 1986, 1988). Thus, one unit of experimentally<br />

determined input (stimulus) energy from the environment can be mapped (c<strong>or</strong>related) inf<strong>or</strong>mationally upon (related to)<br />

one unit of output (<strong>or</strong>ganismal-response) energy using the proposed sulfur-based encoding mechanism. Some key parameters<br />

of <strong>this</strong> experimentally elucidated sulflaydryl / disulfide-dependent electrochemical mechanism are further discussed in Table<br />

2.<br />

Table 2.--Some key fascets of the sulfhydryl / disulfide-dependent electrochemical mechanism involved in insect perception<br />

of phytochemical (e.g. 1,4-naphthoquinone) messengers.<br />

1. Each of the three possible pairings of the chem<strong>or</strong>eception parameters (a, b, e) yields a linear regression with r > 0.95: (a)<br />

the moles of messenger required in a standardized insect behavi<strong>or</strong>al assay to cause a > 99% change in that behavi<strong>or</strong>; (b) the<br />

maximum in vitro polarographic U_/2shift in millivolts by the involved recept<strong>or</strong> and energy-transducer sulflaydryl / disulfide<br />

protein from the insect's antennal chemosens<strong>or</strong>y neurons when saturated with the above messenger: and (e) the maximum<br />

percent inhibition of a standardized excitant-stimulated electroantannogram (EAG) by the above messenger (Rozental and<br />

N<strong>or</strong>ris 1973, 1975; N<strong>or</strong>ris and Chu 1974; N<strong>or</strong>ris 1979, 1986, 1988).<br />

2. Simultaneous solution of the three linear relationships among the parameters in (1) showed that the c<strong>or</strong>relation (rZ= 0.95)<br />

between (a) the maximum U_/2shift elicited in the sulfhydryl / disulfide recept<strong>or</strong> and energy-transducer protein by the<br />

messenger and (b) the maximum percent inhibition of the standardized EAG by that messenger is so high that only one of<br />

these two parameters need be considered in a mathematical description of the transduction of the molar-messenger energy<br />

into insect behavi<strong>or</strong>al change. Log Y - 3.40 - 0.112 Log X, quantifies the energy-transduction relationship (N<strong>or</strong>ris 1986,<br />

11988).<br />

3. Based on behavi<strong>or</strong>al analyses (Rozental and N<strong>or</strong>ris 1975; N<strong>or</strong>ris 1986, 1988), three distinct sets of sulfhydryl (thiol)dependent<br />

recept<strong>or</strong> sites f<strong>or</strong> 2-methyl-1,4-naphthoquinone (menadione) messenger exist in the recept<strong>or</strong> and energy-transducing<br />

protein in Periplaneta americana. Those sulfhydryls in each of these three sets of recept<strong>or</strong> sites cause a 4-5 millivolt<br />

shift in the recept<strong>or</strong> and energy-transducing protein's U_/2value when they react at the mercury electrode involved in <strong>this</strong><br />

polarographic analysis (Rozental and N<strong>or</strong>ris 1973; N<strong>or</strong>ris 1979, 1986, 1988).<br />

4. Based on the EAG-inhibition assay, messenger-menadione saturation of one set of recept<strong>or</strong> sites, as described in (3),<br />

causes about an 8% inhibition (N<strong>or</strong>ris and Chu 1974; N<strong>or</strong>ris 1986, 1988). Thus, the the<strong>or</strong>etical maximal inhibition of the<br />

standardized EAG by saturation of the recept<strong>or</strong>s in all three sites with menadione might be predicted as 3 (sites) times 8%,<br />

which equals 24%. It is interesting and significant that the experimentally determined range in maximal percent of EAG<br />

inhibition by saturation with menadione was 23-25%.<br />

5. Data summarized in (3) and (4) above lead to the interpretation that a recept<strong>or</strong> and energy-transducer Uz/2shift of 4-5 mV<br />

equals an 8% inhibition in the EAG. This means that each 4-5 mV shift in the Uu2 is accompanied by an 8% inhibition in the<br />

standardized EAG. The observed linear relationship between the U_a millivolt shift in the recept<strong>or</strong> and energy-transducer<br />

protein and the percent EAG inhibition shows that the primary encoding of the message dictating the whole-insect behavi<strong>or</strong><br />

occurs in the energy-transducer protein in the chemosens<strong>or</strong>y sensillum (N<strong>or</strong>ris 1979, 1986, 1988).<br />

6. Our research explains f<strong>or</strong> the first time, both in electrochemical and electrophysiological parameters, why the EAG is so<br />

meaningful to an understanding of the chemical senses of insects. EAG does not just measure millivolts of electrical energy,<br />

but also the energy after it has already been coded as inf<strong>or</strong>mation in the recept<strong>or</strong> and energy-transducer protein adequately to<br />

elicit a predictable behavi<strong>or</strong> in (by) the insect (N<strong>or</strong>ris 1979, 1981, 1986, 1988).<br />

7. Saturation of the recept<strong>or</strong> and energy-transducer protein with p-chl<strong>or</strong>omercuribenzoate (PCMB), a compound which reacts<br />

. specifically and irreversibly with sulfhydryls, blocks the above characterized messenger-induced Uu2 shift in the protein.<br />

Thus, the conversion of molar-messenger energy into electrochemically based inf<strong>or</strong>mation adequate to predict whole-insect<br />

behavi<strong>or</strong> is blocked by the sulfhydryl-specific reagent, PCMB (Rozental and N<strong>or</strong>ris 1973; N<strong>or</strong>ris 1979, 1981, t988).<br />

8. Sulfur, as in sulfhydryl / disulfide redox systems in proteins, is the critical dynamic elemental interface between responsive<br />

cells and the stimulating environment, whether biotic <strong>or</strong> abiotic.<br />

53


SUMMARY<br />

Animals and plants "communicate" with each other, and with their environments, via common ion, free radical, and<br />

molecular messengers which function by electrochemical energy-transduction mechanisms. Such mechanisms depend upon<br />

the reversible sulflwdryl-disulfide redox couple in recept<strong>or</strong> and energy-transducer proteins to convert quantitatively messeaget-based<br />

energy states from the environment into altered membrane potentials and/<strong>or</strong> second messengers which may serve<br />

as signals in the elicited cell, and between it and other cells in an <strong>or</strong>ganism. Recept<strong>or</strong> and energy-transducer proteins are<br />

associated with the plasma membrane which surrounds each living cell. Within biological constraints, the conversion of a<br />

molar-messenger energy state into a redox-based energy state in the sulfhydryl-disulfide recept<strong>or</strong> and energy-transducing<br />

protein in the plasma membrane is linear (i.e., quantal). This means that messenger-b<strong>or</strong>ne energy from the environment is<br />

converted to inf<strong>or</strong>mational energy (i.e., units) by the perceiving cell.<br />

Many entomologists and chemical ecologists have used the electroantennogram (i.e., EAG) to detect compounds<br />

from the environment which alter the behavi<strong>or</strong> of insects. In using <strong>this</strong> classical technique, the experimentalist is measuring<br />

change in the energy (e.g., dendritic-membrane potential) state in the primary peripheral chemosens<strong>or</strong>y neurons which are<br />

specially "housed and exposed to the external environment" within the antenna of the insect. The experimental use of the<br />

EAG and the c<strong>or</strong>rect prediction, thereby, of the resultant behavi<strong>or</strong>al change elicited in the whole, live insect constitutes<br />

scientific proof that the inf<strong>or</strong>mation necessary f<strong>or</strong> alteration of the behavi<strong>or</strong> of the whole insect can be encoded in the primary<br />

peripheral chemosensitive neuron. This encoding of energy into inf<strong>or</strong>mation is dependent upon the element 'sulfur', and<br />

especially its readily reversible sulfhydryl (i.e., thiol, -SH) / disulfide (i.e., -S-S-) redox couple. This encodement of chemical-messenger<br />

energy into biologically useful inf<strong>or</strong>mation is blocked (<strong>or</strong> otherwise altered) in the intact cell <strong>or</strong> whole<br />

<strong>or</strong>ganism by the application of biological concentrations of reagents which react specifically with sulfhydryls and/<strong>or</strong> disulfides<br />

in proteins in plasma inembrane. Recent research has proven that <strong>this</strong> transduction of environmental energy into<br />

biologically useful inflmnation also occurs in plant cells. Thus, the sulfhydryl / disulfide-dependent Environmental Energy<br />

Exchange Code is supp<strong>or</strong>ted by extensive scientific findings from both animal and plant reahns.<br />

The unique atomic attributes of the element 'sulfur' f<strong>or</strong> fulfilling <strong>this</strong> vital role in the conversion of environmental<br />

energy into biologically useful inf<strong>or</strong>naation figrall cells were clearly described by Wald (1969). It is f<strong>or</strong>tunate that scientists<br />

can now readily test the role of sulfur, and especially the sulthydryl / disulfide redox couple, in the exchange of environmentally<br />

based energy into inf<strong>or</strong>mation in any living cell. Chemical ecologists seem especially f<strong>or</strong>tunate in <strong>this</strong> regard through<br />

their frequent familiarity with the EAG and other electrophysiological techniques f<strong>or</strong> experimentation. We have also shown<br />

the usefulness of classical electrochemical (e.g., dropping-mercury-electrode polarography) techniques f<strong>or</strong> asking questions<br />

about the roles of sultur and its derivatives in the proposed Environmental Energy Exchange Code. Further experiments on<br />

<strong>this</strong> exciting energy-exchange interface between living cells, <strong>or</strong>ganisms, and their vital environments should yield data which<br />

significantly improve our abilities to quantify environmental influences on the expressions of phenotypes by genomes, and on<br />

the functionalities and longevities of such phenotypes.<br />

ACKNOWLEDGEMENTS<br />

<strong>Research</strong> rep<strong>or</strong>ted here from the auth<strong>or</strong>'s lab<strong>or</strong>at<strong>or</strong>ies was supp<strong>or</strong>ted partially by the College of Agricultural and Life<br />

Sciences, University of Wisconsin, Madison, Wisconsin, USA; in part by numerous research grants from the U.S. National<br />

Science Foundation and the U.S. National Institutes of Health; and in part by U.S. Hatch Projects 3040 and 3419, McIntire-<br />

Stennis Project 3127, and <strong>USDA</strong> Competitive Grants 84-CRCR-1-1501 and 88-37153-4043.<br />

LITERATURE CITED<br />

BELL, E.A. 1981. The physiological role(s) of secondary (natural)products, p. 1-19. In Conn, E.E., ed. The Biochemistry<br />

of Plants: A Comprehensive Treatise. vol 7. Secondary Plant Products. Academic Press, New Y<strong>or</strong>k.<br />

DARVILL, A.G. and ALBERSHEIM, R 1984. Phytoalexins and their elicit<strong>or</strong>s - a defense against microbial infection in<br />

plants. Annu. Rev. Plant Physiol. 35: 243-275.<br />

DIXON, R.A. and LAMB, C.J. 1990. Molecular communication in interactions between plants and microbial pathogens.<br />

54 Annu. Rev. Plant Physiol. Plant Molec. Biol. 41" 339-367.


FRAZIER, J.L. and HEITZ, J.R. 1975. Inhibition ofolfaction in the moth Helio<strong>this</strong> virescens by the sulfhydryl reagent<br />

ftu<strong>or</strong>escein mercuric acetate. Chem. Senses Flav<strong>or</strong> 11:271-281.<br />

HAANSTAD, J.O. and NORRIS, D.M. 1992. Altered elm resistance to smaller European elm bark beetle (Coleoptera:<br />

Scolytidae) and f<strong>or</strong>est tent caterpillar (Lepidoptera: Lasiocampidae). J. Econ. Entomol. 85:172-181.<br />

HAROLD, F.M. 1986. The vital f<strong>or</strong>ce: A study of bioenergetics. W.H. Freeman and Co., San Francisco, 577 p.<br />

KAISSLING, K.E. t971. Insect olfaction, p. 351-431. In Beidler, L.M., ed. Handbook of Sens<strong>or</strong>y Physiology, vol. 5.<br />

Springer-Verlag, Berlin.<br />

KAISSMNG, K.E. 1974. Sens<strong>or</strong>y transduction in insect olfact<strong>or</strong>y recept<strong>or</strong>s, p. 243-273. In Jaenicke, L., ed. Biochemistry<br />

of Sens<strong>or</strong>y Functions. Springer-Verlag, Berlin.<br />

KAISSLING, K.E. 1987. R.W. Wright Lectures on Insect Olfaction, p. 1-75. In Colbert, K., ed. Simon Fraser University,<br />

Burnaby, B.C., Canada.<br />

KOGAN, M. and PAXTON, J. 1983. Natural inducers of plant resistance to insects, p. 153-17 I. In Hedin, RA., ed. Plant<br />

Resistance to Insects. American Chemical Society, Washington, D.C.<br />

LIU, S.H., NORRIS, D.M., HARTWIG, E.E., and XU, M. 1992. Inducible phytoalexins in juvenile soybean genotypes<br />

predict soybean looper resistance in the fully developed plants. Plant Physiol. 100: 1479-1485.<br />

LIU, S.H., NORRIS, D.M., and LI, J. 1994. Peroxidase activity as c<strong>or</strong>related with stress-inducible insect resistance in<br />

Glycine max. Trends in Agricultural Sciences-Entomology, in press.<br />

LIU, S.H., NORRIS, D.M., and XU, M. 1993. Insect resistance and glyceollin concentration in seedling soybeans supp<strong>or</strong>t<br />

resistance ratings of fully developed plants. J. Econ. Entomol. 86: 401-406.<br />

MA, W.C. 1977. Alterations of chem<strong>or</strong>ecept<strong>or</strong> function in armyw<strong>or</strong>m larvae (Spodoptera exempta) by a plant-derived<br />

sesqui-terpenoid and by sulthydryl reagents. Physiol. Entomol. 2: 199-297.<br />

MA, W.C. 1981. Recept<strong>or</strong> membrane function in olfaction and gustation: Implications from modification by reagents and<br />

drugs, p. 267-287. In N<strong>or</strong>ris, D.M., ed. Perception of Behavi<strong>or</strong>al Chemicals. Elsevier/N<strong>or</strong>th-Holland Biomedical<br />

Press, Amsterdam.<br />

MARKOVIC, I., HAANSTAD, J.O., and NORRIS, D.M. 1993. Chemical c<strong>or</strong>relates of alpha-tocopherol (vitamin E) altered<br />

Malacosoma disstria herbiv<strong>or</strong>y in Fraxinus pennsylvanica var. subintegerrinia, green ash. J. Chem. Ecol. 19:1205-<br />

1217.<br />

MORTON, R.A. 1965. Biochemistry of Quinones. Academic Press, London.<br />

NEUPANE, F.R and NORRIS, D.M. 1990. Iodoacetic acid alteration of soybean resistance to the cabbage looper (Lepidoptera:<br />

Noctuidae). Environ. Entomol. 19: 215-221.<br />

NEUPANE, F.R and NORRIS, D.M. 1991a. Sulthydryl reagent alteration of soybean resistance to the cabbage looper,<br />

Trichoplusia ni. Ent. Exp. Appl. 60: 239-245.<br />

NEUPANE, ER and NORRIS, D.M. 1991b. Alpha-tocopherol alteration of soybean anti-herbiv<strong>or</strong>y to Trichoplusia ni larvae.<br />

J. Chem. Ecol. 17: 1941-1951.<br />

NEUPANE, ER and NORRIS, D.M. 1992. Antioxidant alteration of Glycine max (Fabaceae) defensive chemistry: analogy<br />

to herbiv<strong>or</strong>y elicitation. Chemoecology 3: 25-32.<br />

NORRIS, D.M. 1969. Energy transduction mechanism in olfaction and gustation. Nature 222: 1263-1264.<br />

55


NORRIS, DM. 1970. Quinol stimulation and quinone deterrency of gustation by Scolytus multistriatus. Ann. Entomol.<br />

Soc. Amer.63: 476-478.<br />

NORRIS, D.M. 1971. A hypothesized unifying mechanism in neural function. Experientia 27:531-532.<br />

NORRIS, D.M. 1976. How certain insects take the bitter with the sweet. Bull. Entomol. Soc. Amer. 22: 27-30.<br />

NORRIS, D.M. 1979. Chem<strong>or</strong>eception proteins, p. 59-77. In Narahashi, T., ed. Neurotoxicology of Insecticides and Pheromones.<br />

Plenum Publish., New Y<strong>or</strong>k.<br />

NORRIS, D.M. 1981. Possible unifying principles in energy transduction in the chemical senses, p. 289-306. In N<strong>or</strong>ris,<br />

DM., ed. Perception of Behavi<strong>or</strong>al Chemicals. Elsevier/N<strong>or</strong>th-Holland Biomedical Press, Amsterdam.<br />

NORRIS, DM. 1985. Electrochemical parameters of energy transduction between repellent naphthoquinones and lipoprotein<br />

recept<strong>or</strong>s in insect neurons. Bioelectrochem. Bioenerget. 14: 449-456.<br />

NORRIS, DM. 1986. Anti-feeding compounds, p. 97-146. In Haug, G. and Hoffmann, H., eds. Chemistry of Plant Protection,<br />

vol. 1. Springer-Verlag, Heidelberg.<br />

NORRIS, DM. 1988. Periplaneta americana perception of phytochemical naphthoquinones as allelochemicals. J. Chem.<br />

Ecol. 14: 1807-1819.<br />

NORRIS, D.M. 1994. Phytochemicals as messengers altering behavi<strong>or</strong>. In Ananthakrishnan, T.N., ed. Functional Dynamics<br />

of Phytophagous Insects. Oxf<strong>or</strong>d and IBH Publish. Co. Pvt. Ltd., New Delhi, India, in press.<br />

NORRIS,D.M., BAKER, J.E., BORG, T.K., FERKOVICH, S.M., and ROZENTAL, J.M. 1970a. An energy-transduction<br />

mechanism in chem<strong>or</strong>eception by the bark beetle, Scolytus multistriatus. Contrib. Boyce Thomp. Inst. 24: 263-274.<br />

NORRIS, D.M. and CHU, H.M. 1974. Chemosens<strong>or</strong>y mechanism in Periplaneta americana: Electroantennogram comparisons<br />

of certain quinone feeding inhibit<strong>or</strong>s. J. Insect Physiol. 20: 1687-1696.<br />

NORRIS, D.M., FERKOVICH, S.M., BAKER, J.E., ROZENTAL, J.M., and BORG, T.K. 1971. Energy transduction in<br />

quinone inhibition of insect feeding. J. Insect Physiol. 17: 85-97.<br />

NORRIS, D.M., FERKOVICH, S.M., ROZENTAL, J.M., BAKER, J.E., and T.K. BORG. 1970b. Energy transduction:<br />

inhibition of cockroach feeding by naphthoquinones. Science 170: 754-755.<br />

NORRIS, DM. and LIU, S.H. 1992. A common chemical mechanism f<strong>or</strong> insect-plant communication, p. 186-187. In<br />

Menken, S.B.J., Visser, J.H., and Harrewijn, P., eds. Proceedings of 8th International Symposium on Insect-Plant<br />

Relationships. KluwerAcademic Publisher, D<strong>or</strong>drecht, The Netherlands.<br />

O'CONNELL, R.J. 1981. The encoding of behavi<strong>or</strong>ally imp<strong>or</strong>tant od<strong>or</strong>ants by insect chemosens<strong>or</strong>y recept<strong>or</strong> neurons, p.<br />

133-163. In N<strong>or</strong>ris, D.M., ed. Perception of Behavi<strong>or</strong>al Chemicals. Elsevier/N<strong>or</strong>th-Holland Biomedical Press,<br />

ATnsterdam.<br />

PACKER, L. and GLAZER, A.N., eds. 1990. Oxygen radicals in biological systems, part B, Oxygen radicals and antioxidants,<br />

p. 1-855. In Abelson, J.N. and Simon, M.I., eds.-in- chief. Methods in Enzymology. vol. 186. Academic Press<br />

New Y<strong>or</strong>k.<br />

PYLE, A.M. 1993. Ribozymes: a distinct class of metalloenzymes. Science 261:709-714.<br />

RAINA, A.K., KINGAN, T.G., and MATTOO, A.K. 1992. Chemical signals from host plant and sexual behavi<strong>or</strong> in a naoth.<br />

Science 255: 592-594.<br />

RAVICHANDRAN, K.G., BODDUPALLI, S.S., HASEMANN, C.A., PETERSON, J.A., and DEISENHOFER, J. 1993.<br />

56 Crystal structure of hemoprotein domain of P450BM-3, a prototype f<strong>or</strong> microsomal P450's. Science 261" 731-7 36.


REICHARD, R 1993. From RNA to DNA, why so many ribonucleotide reductases? Science 260: 1773-1777.<br />

RODRIGUEZ, E. and LEVIN, D.A. 1976. Biochemical parallelisms of repellents and attractants in higher plants and<br />

arthropods, p. 214-270. In Wallace, J.W. and Mansell, R.L., eds. Biochemical Interaction Between Plants and Insects.<br />

Rec. Adv. Phytochem. I0: 214-270.<br />

ROZENTAL, J.M. and NORRIS, D.M. 1973. Chemosens<strong>or</strong>y mechanism in American cockroach olfaction and gustation.<br />

Nature 244: 370-371.<br />

ROZENTAL, J.M. and NORRIS, D.M. 1975. Genetically variable olfact<strong>or</strong>y recept<strong>or</strong> sensitivity in Periplaneta americana.<br />

Life Sciences 17:105-110.<br />

RYAN, C.A. 1983. Insect-induced chemical signals regulating natural plant protection responses, p. 43-60. In Denno, R.F.<br />

and McClure, M.S., eds. Variable Plants and Herbiv<strong>or</strong>es in Natural and Managed Systems. Academic Press, New<br />

Y<strong>or</strong>k.<br />

SELA, M., WHITE, F.H., and ANFINSEN, C.B. 1957. Reductive cleavage of disulfide bridges in ribonuclease. Science<br />

125: 691-692.<br />

SEQUEIRA, L. 1983. Mechanisms of induced resistance in plants. Annu. Rev. Microbiol. 37: 51-79.<br />

SHARMA, H.C., and NORRIS, D.M. 1991. Chemical basis of resistance in soya bean to cabbage looper, Trichoplusia hi. J.<br />

Sci. Food Agric. 55: 353-364.<br />

SLINGER,G., ROZENTAL, J.M., and NORRIS, l).M. 1975. Sulfhydryl groups and the quinone recept<strong>or</strong> in insect olfaction<br />

and gustation. Nature 256: 222-223.<br />

STAMLER, J.S., SINGEL, D.J., and LOSCALZO, J. 1992. Biochemistry of nitric oxide and its redox-active f<strong>or</strong>ms. Science<br />

258: 1898-1902.<br />

STORZ, G., TARTAGLIA, L.A., and AMES, B.N. 1990. Transcriptional regulat<strong>or</strong> of oxidative stress-inducible genes: direct<br />

activation by oxidation. Science 248: 189-194.<br />

STOSSEL, A. 1984. Regulation by sulfhydryl groups of glyceollin accumulation in soybean hypocotyls. Planta 160: 314-<br />

319.<br />

SZENT-GYORGYI, A. 1973. The development of bioenergetics, p. 1-4. In Avery, J., ed. Membrane Structure and Mechanisms<br />

of Biological Energy Transduction. Plenum Press, New Y<strong>or</strong>k.<br />

TEMPLETON, M.D. and LAMB, C.J. 1988. Elicit<strong>or</strong>s and defense gene activation. Plant Cell Environ. 11: 395-401.<br />

VANl)E BERG, J.S. 1981. Ultrastructural and cytochemical parameters of chemical perception, p. 103-131. In N<strong>or</strong>ris,<br />

l).M., ed. Perception of Behavi<strong>or</strong>al Chemicals. Elsevier/N<strong>or</strong>th-Holland Biomedical Press, Amsterdam.<br />

VAN DER VLIET, A. and BAST, A. 1992. Effect of oxidative stress on recept<strong>or</strong>s and signal transmission. Chem. Biol.<br />

Interact. 85:95-116.<br />

VILLET, R.H. 1974. Involvement of amino and sulfhydryl groups in olfact<strong>or</strong>y transduction in silk moths. Nature 248: 707-<br />

708.<br />

VOGT, R.G. and RIDDIFORD, L.M. 1986. Pheromone reception: a kinetic equilibrium, p. 201-208. In Payne, T.L., Birch,<br />

M.C., and Kennedy, C.E.J., eds. Mechanisms in Insect Olfaction. Clarendon Press, Oxf<strong>or</strong>d, U.K.<br />

WALl), G. 1969. Life in the second and third periods; <strong>or</strong> why phosph<strong>or</strong>us and sulfur f<strong>or</strong> high-energy bonds?, p. 156-t67. In<br />

Kalckar, H.M., ed. Biological Phosph<strong>or</strong>ylations: Development of Concepts. Prentice-Hall, New Y<strong>or</strong>k.<br />

57


FUNGAL ENDOPHYTES: CONTRASTING EFFECTS IN<br />

TREES AND GRASSES<br />

STANLEY H. FAETH<br />

Department of Zoology, Arizona State University, Tempe, Arizona 85287-1501, USA<br />

INTRODUCTION<br />

Fungal endophytes, fungi that live asymptomatically and intercellularly within tissues of most plants, have recently<br />

received increasing attention from ecologists (e.g., Strong 1988). The presence of fungal endophytes in plants, especially<br />

grasses, can increase resistance to herbiv<strong>or</strong>es due to the production of alkaloidal mycotoxins, but also may increase resistance<br />

to drought and flooding stress, increase plant competitive abilities, and, f<strong>or</strong> endophytes transmitted vertically via seeds, deter<br />

seed predat<strong>or</strong>s <strong>or</strong> increase seed dispersal (Clay 1990, Knoch et al. 1993). Endophytic fungi are typically considered plant<br />

mutualists because of these potential modes of increased plant fitness (Clay 1990). However, most research on fungal<br />

endophytes has involved introduced, agricultural f<strong>or</strong>age grasses, such as tall rescue <strong>or</strong> perennial ryegrass. Little is known of<br />

the ecological role of endophytes in either native grasses <strong>or</strong> woody plants relative to plant-herbiv<strong>or</strong>e, plant-plant, <strong>or</strong> plantseed<br />

predat<strong>or</strong> interactions. Here, I summarize our research to date on endophytes in Quercus em<strong>or</strong>yi (Em<strong>or</strong>y oak) and their<br />

effects on a maj<strong>or</strong> herbiv<strong>or</strong>e, the leafminer Cameraria sp. nov. (Lepidoptera:Gracillaridae). Leafminer larvae spend 11<br />

months within leaf tissues and are confined to a single leaf chosen by the ovipositing female. Theref<strong>or</strong>e, endophytes should<br />

alter leafininer perf<strong>or</strong>mance m<strong>or</strong>e than that of mobile, exophytic insects. I contrast these results with endophytes in Festuca<br />

arizonica (Arizona fescue). I predict stronger effects of endophytes on herbiv<strong>or</strong>es and seed predat<strong>or</strong>s of Arizona fescue due<br />

to differences in mode of fungal transmission and specificity,<br />

i_:<br />

METHODS ::<br />

We have monit<strong>or</strong>ed seasonal and spatial patterns of fungal endophyte infections and the leafminer in trees of Em<strong>or</strong>y :i<br />

oak at Oak Flat study area in central Arizona f<strong>or</strong> the past 4 and 10 years (Faeth 1991), respectively. We have isolated at least<br />

12 species of endophytes from Em<strong>or</strong>y oak, but four species, QE1 (Asteromella sp.), QE2 (Ascomycete:Diap<strong>or</strong>thales), QE7<br />

(Plecophomella sp.), and Y1 (filamentous yeast) make up >95% of all infections. All of these endophytes are transmitted<br />

h<strong>or</strong>izontally via sp<strong>or</strong>es, likely carried in rainsplash. In observational studies, we have c<strong>or</strong>related the presence of living and<br />

dead larval leafminers with the frequency of infection. In manipulative experiments, we have either increased (sp<strong>or</strong>e<br />

spraying of leaves <strong>or</strong> sp<strong>or</strong>e injection of individual mines) <strong>or</strong> decreased (enclosing branches with plastic <strong>or</strong> application of<br />

fungicides) to test the role of individual fungal endophytes on leafminer developmental and m<strong>or</strong>tality.<br />

We have begun to investigate the role of fungal endophytes in Arizona rescue populations and their relationship to<br />

intensity of grazing andsoil nutrients. Arizona rescue harb<strong>or</strong>s two endophytes, Acremonium starrii and a p-endophyte<br />

(Phialoph<strong>or</strong>a-like). Both endophytes are transmitted vertically from maternal to offspring plant via seed, but the p-endophyte<br />

also sp<strong>or</strong>ulates and can be transmitted h<strong>or</strong>izontally fromadult plant to adult plant. We have conducted preliminary<br />

experiments testing the role of the Acremonium endophyte in reducing seed predation and increasing seed dispersal by seed<br />

harvesting ants (Pogonomymrex species) by presenting E+ (infected) and E- (uninfected) rescue seeds to ant colonies and<br />

following prop<strong>or</strong>tions collected and discarded into refuse piles.<br />

RESULTS AND DISCUSSION<br />

Generally, endophyte infections in oaks increase seasonally, with the peak infection level coinciding with summer<br />

rains in Arizona in July-August. Overall infection levels vary with all spatial scales- between localities, between and within<br />

Mattson, W.J., Niemel_i, p., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. LISDA<br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN<br />

58


;<br />

trees, and between and within individual leaves. Differences in infection levels between trees result from dif{Erences in<br />

infection by individual species, with QE7 comprising differences early in the season and QE1 later in the growing season.<br />

Leafmining is associated with increased fungal infection, and leaves with dead larvae have higher infections than those with<br />

live larvae. However, neither mass n<strong>or</strong> survival of leafminers was affected by sp<strong>or</strong>e-injection of the three common endophytes.<br />

One endophyte (Y 1) increased developmental thne. Other experiments show that oviposition is not associated with<br />

leaves that are either' m<strong>or</strong>e <strong>or</strong> less likely to become infected with endophytes. Thus, leafmining activity appears to increase<br />

infection frequency, probably by altering the surface of the leaf, but the endophytes themselves appear to have only weak<br />

effects on leafminer development and survival.<br />

Preliminary observational evidence :indicates that Arizona fescue in areas under intensive grazing by cattle and native<br />

ungulates have higher frequency of Ac'rernonium starrii but not the p-endophyte. These results supp<strong>or</strong>t our prediction that<br />

endophytes that are only transmitted vertically are m<strong>or</strong>e likely to interact mutualistically with the plant and provide protection<br />

against herbiv<strong>or</strong>es. Fur<strong>or</strong>e experiments will test the dependency of the mutualism on available resources to the grass and<br />

controlled levels of herbiv<strong>or</strong>y. Similarly, experiments with seed-h_vesting ants show that Acremoni_ma-infected seeds are<br />

less likely to be collected. Of the seeds that are harvested, E+ seeds are m<strong>or</strong>e likely than E- seeds to be discarded into reff_se<br />

piles where germination success is higher than in surrounding areas.<br />

SUMMARY<br />

Endophytic fungi are diverse and ubiquitous in almost all woody and non-woody plants examined to date. The role<br />

of endophytes in plant-herbiv<strong>or</strong>e, plant-plant, and plant-seed predat<strong>or</strong> interactions in natural systems is still largely unex-<br />

p]<strong>or</strong>ed. Testable hypotheses and predictions, however, can be made about the direction and strength of the interactions based<br />

upon mode of transmission and specificity of the fungi and plant.<br />

ACKNOWLEDGMENTS<br />

I thank Kyle Hammon and Dennis Wilson f<strong>or</strong> preliminary data and Rindy Anderson, Carmen Febus, Dawn<br />

Hagerman, Catherine Hudson, Terri Kingrey, Myrna Miller, and Franziska Schulthess f<strong>or</strong> technical assistance. This w<strong>or</strong>k is<br />

supp<strong>or</strong>ted by NSF Grant BSR 91-07296 to SHE<br />

LITERATURE CITED<br />

CLAY, K. 1990.. Fungal endophytes of grasses. Ann. Rev. Ecol. Syst. 21: 275-297.<br />

FAETH, S.H. 199t. Effect of oak leaf size on abundance, dispersion, and survival of the leafminer, Cameraria sp. nov.<br />

(Lepidoptera:Gracillariidae). Envir. Ent. 20: 196-204.<br />

KNOCH, T.R., I_ETH, S.H., and ARNOTT, D.L. 1993. Endophytic fungi alter f<strong>or</strong>aging and dispersal by desert seed harvesting<br />

ants. Oecologia 95: 470-475.<br />

STRONG, D.R. 1988. Endophytic mutualism and plant protection from herbiv<strong>or</strong>es. Ecol. 69: 1.<br />

59


PROPAGATING THE EFFECTS OF HERBIVORE ATTACK IN AN<br />

OBJECT-ORIENTED MODEL OF A TREE<br />

JARI PERTTUNEN, HANNU SAARENMAA, RISTO SIEV,_NEN, HANNU SALMINEN,<br />

ANTTI POUTTU, and JOUNI V,_KEV,_<br />

Finnish F<strong>or</strong>est <strong>Research</strong> Institute, Unioninkatu 40A, SF-00170, Helsinki, Finland<br />

INTRODUCTION<br />

Understanding cause and effect in f<strong>or</strong>est health problems requires an understanding of the structure and functioning<br />

of trees. Bef<strong>or</strong>e conclusions about a malfunctioning system can be made, we must first understand how a healthy one w<strong>or</strong>ks,<br />

In industrial engineering and medicine, diagnostic reasoning about structure have long traditions (e.g., Bratko et al. 1989,<br />

T<strong>or</strong>asso and Console 1989), and well defined structural models exist f<strong>or</strong> man-made machines and the human body. Diagnosis<br />

of dis<strong>or</strong>ders in trees and f<strong>or</strong>ests is not that advanced. One of the reasons f<strong>or</strong> <strong>this</strong> is that we have not had good structural<br />

models of trees at our disposal.<br />

By structural models, we mean models that make the parts of the system explicit. Mathematical models describing<br />

photosynthesis, respiration, water pressure, etc. are not enough alone. They have to be bound into some larger structure, like<br />

shoots, buds, needles, etc. These parts in turn must be explicitly connected with their physical immediate neighb<strong>or</strong>s. Such a<br />

model of a tree would f<strong>or</strong>m a topology of thousands of interconnected parts. L-systems (Prunsinkiewicz and Hanan 1989)<br />

and fractal trees (Mandelbrot 1977) are probably the <strong>or</strong>iginal approaches to <strong>this</strong> problem. However, those models are made<br />

f<strong>or</strong> visualization, and do not maintain the status of the parts after they have been graphically drawn. Plant and tree growth<br />

has been studied in detail f<strong>or</strong> research in plant m<strong>or</strong>phology and landscape architecture (e.g., Reffye et al. 1989). However,<br />

structural models of tree development and survival have emerged only lately because the necessary software technologies<br />

have only been in place f<strong>or</strong> a few years. Sequeira et al. (1991) have presented an structural model of plant. Ahonen and<br />

Saarenmaa (1991) and Saarenmaa (1991) have prototyped a topological tree. Nikinmaa (1992) has a presented the models<br />

f<strong>or</strong> the functioning of a structurally explicit Scots pine tree.<br />

Object-<strong>or</strong>iented (OO) programming is rapidly becoming the leading way of constructing software. OO software<br />

consists of objects that are m<strong>or</strong>e <strong>or</strong> less encapsulated entities. They combine attributes and behavi<strong>or</strong> under the same encapsulation.<br />

Objects can consist of other objects and they can communicate with each other with messages. They make it possible<br />

to represent structures such as biological <strong>or</strong>ganisms with a great realism. Once an object-<strong>or</strong>iented tree model has been built,<br />

it can, hopefully, be reused f<strong>or</strong> many different purposes. These uses may be growth prediction and physiological research.<br />

Such models should also be useful f<strong>or</strong> making diagnoses and visualizing the effects of dis<strong>or</strong>ders. Herbiv<strong>or</strong>e-tree interactions<br />

are often so complicated that they can only be understood through modelling. An OO tree model should be able to propagate<br />

the effects of a herbiv<strong>or</strong>e attack in one part of the tree into others where the symptoms and secondary damage may be<br />

observed.<br />

In <strong>this</strong> paper we describe briefly an OO model of Scots pine, Pinus sylvestris L., that we have built, and demonstrate<br />

its uses f<strong>or</strong> simulating dis<strong>or</strong>ders caused by two herbiv<strong>or</strong>es,<br />

METHODS<br />

Object-<strong>or</strong>iented modelling allows building software that mechanistically resembles the real w<strong>or</strong>ld. This similarity is<br />

achieved with three basic concepts: encapsulated objects, classification of objects, and message passing between objects<br />

(Rumbaugh et al. 1991, Saarenmaa et al. 1994).<br />

Mattson, W.J., Niemel_i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. LISIDA<br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

60


Encapsulated objects are data structures that always maintain their identity, have a clear boundary, and show to the<br />

outside w<strong>or</strong>ld only selected facets of themselves. Objects have attributes and behavi<strong>or</strong>. They can be concrete physical<br />

things, abstract ones, <strong>or</strong> events. Examples of concrete things are trees, insects, vehicles, people, etc. Examples of abstract<br />

objects are beliefs, statistical inf<strong>or</strong>mation, and time. Examples of events are decisions, insect outbreaks, treatments, etc. The<br />

attributes contain the internal state of objects in their values. Examples of the attributes of an insect object are length, wing<br />

col<strong>or</strong>, etc. Examples of behavi<strong>or</strong>s (also called methods) of in an insect object are development, emergence, oviposition, etc.<br />

Objects fall into two categ<strong>or</strong>ies: classes and instances. A class is a definition, a blue<strong>print</strong> f<strong>or</strong> actual instances. F<strong>or</strong><br />

instance, a biological taxon such as the species of "Scots pine" can be a class. Its instances, in turn, are the physical individual<br />

trees that grow outside <strong>this</strong> building. Classes often f<strong>or</strong>m hierarchies where subclasses inherit the attributes and<br />

methods of their superclasses. At a lower level, specialized values and new attributes and methods can be introduced in <strong>or</strong>der<br />

to refine the generic superclasses.<br />

Objects communicate by sending messages to each other. Different objects can respond to the same message in their<br />

own ways. F<strong>or</strong> instance, a "time" object may send a message of "lightness change" with a value "dusk" to all the objects in a<br />

simulation. One insect object may respond to it by activating its flight behavi<strong>or</strong> whereas another, that would only fly in<br />

sunlight, would deactivate its own.<br />

Objects may consist of other objects. F<strong>or</strong> instance, a tree object consists of a root system, one <strong>or</strong> several trunks, and a<br />

crown, which, in turn, consist of smaller objects. Objects may be associated and connected with other objects by having their<br />

names f<strong>or</strong> some purpose as the value of an attribute. OO models should be implemented with an OO programming language.<br />

The present system, called Lignum, has been written in C++ under the Unix operating system, X-Window System, and 3dimensional<br />

interactive graphics.<br />

RESULTS<br />

Description of the Tree Model<br />

The model describes a single Scots pine tree that consists of small elementary parts. The functions of the model have<br />

been further elab<strong>or</strong>ated from Nikinmaa (1992), and the model has been described in detail by Perttunen et al. (1994) and<br />

Salminen et al. (1994). The parts belong currently only into two operational classes: shoots and buds. In addition, we define<br />

classes f<strong>or</strong> the stem, branch wh<strong>or</strong>ls, foliage, bark, phloem, sapwood, and heartwood. Each instance of a stem contains zero <strong>or</strong><br />

m<strong>or</strong>e instances of shoot, branch wh<strong>or</strong>l, and it ends with an instance of bud. Each pair of instances of shoot is separated by an<br />

instance of branch wh<strong>or</strong>l (Fig. 1). Shoots consist of foliage, bark, phloem, sapwood, and heartwood. In <strong>this</strong> version of the<br />

model, the objects in a stem are connected to each other using a list structure.<br />

The tree is <strong>or</strong>iginally started from one bud instance. Buds have methods f<strong>or</strong> creating shoot instances, branch wh<strong>or</strong>l<br />

instances, and new bud instances. The shoots create the parts they consist of. Shoots and branch wh<strong>or</strong>ls that do not have<br />

foliage any longer, are collapsed into the stems.<br />

As we do not yet have a class f<strong>or</strong> foliage, the implementation of shoot captures the productive metabolism of the tree.<br />

These include methods f<strong>or</strong> photosynthesis, respiration, and growth. Foliage is implemented as an attribute f<strong>or</strong> the needle<br />

mass, which must have a value greater than zero f<strong>or</strong> any of the production to take place. Production is handled by a method<br />

which boasts the following function: where Wf is the foliage mass of the shoot and i the degree of interaction of shoot p<br />

which is dependent on the shadiness of the location (Sievanen 1992) with the foliage mass x above the height z. Production<br />

has to be multiplied with fact<strong>or</strong> P0 which describes the photosynthesis under unshaded conditions (Nikinmaa 1992).<br />

Respiration of buds, foliage, phloem, and sapwood is directly prop<strong>or</strong>tional to their mass. In the current implementation,<br />

the root system of the tree is extremely simple. It is considered to have mass that increases as the result of the production<br />

in the shoots. Partof the root mass dies annually. It also respires similarly to the foliage and the sapwood.<br />

After production and respiration the tree can allocate its net production into the new shoots, foliage and roots. The<br />

equations f<strong>or</strong> these are given by Perttunen et al. (1994). Nutritional status of the parts is only dealt with a single attribute in<br />

the present model.<br />

61


Figure I.--Schematic representation of the object tree. (1)<br />

® Bud<br />

F--q Shoot<br />

_


_re3.--The Lignum model can be controlled through windows and sliders and the parameters can be customized. In the<br />

graph shown, one shoot in the top wh<strong>or</strong>l has turned brown due to damage in sapwood as caused by Tomicus spp.<br />

Examples of Herbiv<strong>or</strong>e Attack on the Model ]¥ee<br />

Propagation of the effects of herbiv<strong>or</strong>e attack in the tree model can only be implemented through material flows from<br />

object to another. So far, we have only implemented an upward water flow-method f<strong>or</strong> the sapwood part of shoots and<br />

ds and a downward starch flow-method f<strong>or</strong> the phloem part of shoots and wh<strong>or</strong>ls. The status-attribute of these parts can<br />

values between 0 and 1. The rate of the flow is directly prop<strong>or</strong>tional to <strong>this</strong> status. The status of phloem, sapwood, and<br />

Lgeis also shown graphically as the col<strong>or</strong> of the part as Figure 4 shows.<br />

Sapwood.Status _.- Sapwood.Upflow<br />

,. + +_,,'"_+<br />

".X BLUE Foliage.Status<br />

Shoot.Colour + _ + ..... GREEN f.,,,,,,,.,,._..-,--<br />

+<br />

, /<br />

RED - :_<br />

/<br />

Foliage.Production<br />

/ +<br />

Phloem.Status _,. Phloern.Downflow +<br />

1re 4.--The col<strong>or</strong> of the diseased shoot is shown as the combination of three main status attributes of its parts.<br />

63


B<strong>or</strong>ing by Pine Shoot Beetles<br />

Examples of insects feeding on the canopies of trees are the pine shoot beetles, Tomicus spp. This bark beetle is the<br />

first one to invade dying <strong>or</strong> felled pines to breed in spring. The progeny emerging from breeding sites in trunks disperse to<br />

the crowns of pines during mid- and late-summer. In <strong>or</strong>der to become mature, young adults feed in pine shoots burrowing<br />

intothe pith. The beetles attack mainly current year shoots of upper half of the canopy. The burrows may heal over only in<br />

the thickest shoots, otherwise they turn brown, break, and fall down. Physiologically, Tomicus attacks in shoots have an<br />

effect like the sudden cutting of ducts (Fig. 5).<br />

%<br />

f:: ' • _ _: ,:,. . Figure 5. After about 10 years of repeated Tomicus piniperda damage the tops of<br />

• _- _ i,,__. 5 Scots pine have become def<strong>or</strong>med as on the right. Under a permanent<br />

__ ': '*'.... " " beetle pressure the spike-like tops die and the pines start looking flat-<br />

_.w-_ "_'g_ ;:,:_ ,;. ,.(._ '. _:::':: i _: topped.<br />

, • _,: ":<br />

The effects of removal of shoots on remaining parts of the tree can be divided into three groups: first, the environmental<br />

effects, e.g., the changes in light conditions, water availability and release of nutrients from fallen shoots; second, the<br />

structural reactions like leader shoot changes, f<strong>or</strong>mation of adventitious shoots, sapwood-heartwood relation, and the dying of<br />

roots as the pipe model indicates; and third, the carbohydrate dynamics of the tree, including changes in assimilation<br />

nutrient allocation, root activity and compensation processes. The last group of effects is the most difficult to include<br />

model. These effects are discussed th<strong>or</strong>oughly by Swedish researchers (Fagerstrom et al. 1977, Ericsson et al. 1985,<br />

Langstrom et al. 1990,Troeng and Langstrom 1991, Langstrom and Hellqvist 1991, 1992). In addition, the reactions<br />

capacity,<br />

in a<br />

seem to<br />

d<br />

m<br />

b<br />

o<br />

be different acc<strong>or</strong>ding to age, growing site, and provenance of trees, v<br />

t<br />

The distribution of damage classes of crown can be used to make rough estimates of growth losses in stands repeat-<br />

edly attacked by pine shoot beetles (Kukkola et al. 1994). With a stand of model trees, the spatial pattern of beetle dispersal<br />

and the distribution of growth losses of pines may be examined, b<br />

The present model should be able to simulate the pruning pattern of Tomicus attack and the growth loss. The model<br />

o<br />

t<br />

is currently being used <strong>this</strong> way by letting individual Tomicus object instances to choose a shoot f<strong>or</strong> f<strong>or</strong>aging. The choices<br />

attacking beetles depend on the shoots already allocate by the previously attacked beetles, and the amount the tree stands<br />

above <strong>or</strong> below its neighb<strong>or</strong>s.<br />

of<br />

out<br />

r<br />

Defoliation by Insect Larvae<br />

Scots pine is adapted to low nutrient resources and st<strong>or</strong>es its maj<strong>or</strong> reserves in foliage. Figure 6 shows the dynamics e<br />

that follow from a severe defoliation in <strong>this</strong> situation. A severe defoliation reduces the carbohydrates and nutrients overall in o<br />

the tree, and also reduces the needle number. This leads to increased nutrient concentration in the remaining needles, as the s<br />

fine root biomass is not immediately reduced and its nutrient uptake pumps m<strong>or</strong>e nutrients into the remaining needles. This a<br />

increases the quality of the remaining needles f<strong>or</strong> food of the defoliat<strong>or</strong>. Increased nutrition in needles also leads to their d<br />

increased size and increases production. The carbohydrate concentration in needles increases as well as the concentration of<br />

the carbon-based protective metabolites (f<strong>or</strong> a review, see Herms and Mattson 1992).<br />

64<br />

F<br />

d


SEVERE DEFOLIATION<br />

Tree<br />

Fine Root<br />

+<br />

_<br />

/_'<br />

/<br />

........................ _ Carbohydrate<br />

Reserves<br />

Biomass - J_ Photosynthesis 14Y'+<br />

Nutrient +<br />

Res erv es<br />

1<br />

+<br />

Tree<br />

ee Number<br />

/ .,._.._<br />

Carbohydrate<br />

Uptake Concentration<br />

Needle Size<br />

Capacity 'X_+ Concentration Needle Nutrient<br />

I +<br />

Nutrient | Needle C-Based<br />

Abs<strong>or</strong>ption , + Secondary<br />

Needle Food Quality _ Metabolites<br />

(No Induction) -<br />

Figure 6.--Dynamics of nutrients and carbohydrates in the Lignum model, with the effects of"defoliation.<br />

The present model, as it has explicit descriptions f<strong>or</strong> foliage in each shoot, is capable of duplicating <strong>this</strong> kind of<br />

dynamics. However, we are still calibrating the model, and can not yet claim err<strong>or</strong>-free predictions.<br />

DISCUSSION<br />

The present model was created f<strong>or</strong> a testbench of ideas in many different disciplines. It should be reusable f<strong>or</strong><br />

diagnostic reasoning, studying the effects of herbiv<strong>or</strong>e attack, visualization of symptoms, research on growth and development,<br />

and modelling tree architecture. Although we are still w<strong>or</strong>king on the first version of the model, the experiences have<br />

been promising. Future versions will be built to inc<strong>or</strong>p<strong>or</strong>ate simulation on parallel computer hardware, and zooming in and<br />

out of the tree. The col<strong>or</strong> schemes that we use f<strong>or</strong> illustrating the status of the functions of the tree do not produce realistic<br />

visualization of the symptoms as of now. We will study the col<strong>or</strong> effects that follow from various injuries, and try to animate<br />

the suffering and death of trees.<br />

The model is a part of a larger system f<strong>or</strong> integrated f<strong>or</strong>est health management (Saarenmaa et al. 1994). It has been<br />

built compatible with object-<strong>or</strong>iented descriptions of f<strong>or</strong>est insects and fungi. These have attributes that describe what parts<br />

of trees, represented as object classes, they attack and what kind of disturbances they cause in those parts. Matching OO<br />

trees with OO insects and other harmful agents will create new kind of ecological simulations and also be useful in diagnostic<br />

reasoning.<br />

SUMMARY<br />

A model of a tree which consists of thousands of small interconnected parts can potentially be useful f<strong>or</strong> studying the<br />

effects of insect and disease attack on the whole system. We present such a model f<strong>or</strong> Pinus sylvestris, which is based on<br />

object-<strong>or</strong>iented programming. The model consists of classes f<strong>or</strong> shoots, buds, the stem, branch wh<strong>or</strong>ls, foliage, bark, phloem,<br />

sapwood, and heartwood. The model simulates primary production, respiration, growth, water flow, and starch flow, and has<br />

a geometry that can be monit<strong>or</strong>ed in 3-D col<strong>or</strong> on computer screen. The effects of attacks by shoot b<strong>or</strong>ing bark beetles and<br />

defoliating insects are studied.<br />

65


LITERATURE CITED<br />

AHONEN, J. and SAARENMAA, H. 1991. Model-based reasoning about natural ecosystems: An alg<strong>or</strong>ithm to reduce the<br />

computational burden associated with simulating multiple biotic agents, 193-200. In Computer Science f<strong>or</strong> Environmental<br />

Protection. Sixth Symposium. Munich, Germany, December 4-6, 1991. Springer-Verlag, Heidelberg.<br />

BRATKO, I., MOZETIC, I., and LAVRAC, N. 1989. Kardio. A study in deep and qualitative knowledge f<strong>or</strong> expert systems.<br />

260 p. The MIT Press, Cambridge, Massachusetts.<br />

ERICSSON, A., HELLQVIST, B., LANGSTROM, B., LARSSON, S., and TENOW, O_ 1985. Effects on grow*h of simulated<br />

and induced shoot pruning by Tomicus piniperda as related to carbohydrate and nitrogen dynamics in scots pine.<br />

LAppl. Ecol. 22: 105-124.<br />

•FAGERSTROM, T., LARSSON, S., LOHM, U., and TENOW. O. 1977. Growth in scots pine (Pinus sitvestris L.): a<br />

hypothesis on response to Blastophagus piniperda L. (Col., Scolytidae) attacks. F<strong>or</strong>. Ecol. and Manage. 1:273-281.<br />

HERMS, D.A. and MATTSON, W.J. 1992. The dilemma of plants: to grow <strong>or</strong> defend. The Quart. Rev. Biol. 67(3): 283-<br />

335.<br />

KUKKOLA, M., ANNILA, E., TIMONEN, M., and VARAMA, M. 1994. The development and yield of the Scots pine<br />

stand repeatedly attacked by the pine shoot beetles. Manuscript. Finnish F<strong>or</strong>est <strong>Research</strong> Institute.<br />

KURTH, W. 1992. M<strong>or</strong>phological models of plant growth: Possibilities and ecological relevance. 14 p. ISEM's 8th<br />

International Conference on the State-of-the-Art in Ecological Modelling, Kiel, Germany, September 28 - October 2,<br />

1992.<br />

LANGSTROM, B. and HELLQVIST, C. 1991. Effects of different pruning regimes on growth and sapwood area of Scots<br />

pine. F<strong>or</strong>. Ecol. and Manage. 44: 239-254.<br />

LANGSTROM, B. and HELLQVIST, C. 1992. Height growth recovery and crown development in top-damaged Pinus<br />

sylvestris trees. Scandinavian J. F<strong>or</strong>. Res. 7: 237-247.<br />

L,_NGSTROM, B., TENOW, O., ERICSSON, A., HELLQVIST, C., and LARSSON, S. 1990. Effects of shoot pruning on<br />

stem growth, needle biomass, and dynamics of carbohydrates and nitrogen in Scots pine as related to season and tree<br />

age. Can. J. F<strong>or</strong>. Res. 20: 514-523.<br />

LINDENMEYER, A, 1968. Mathematical models f<strong>or</strong> cellular interaction in development. Parts I and II. J. The<strong>or</strong>et. Biol.<br />

18: 280-315.<br />

MANDELBROT, B.B. 1977. The fractal geometry of nature. 468 p. W.H. Freeman and Co., New Y<strong>or</strong>k.<br />

NIKINMAA, E. 1992. Analyses of the growth of Scots pine; matching structure with functions. Acta F<strong>or</strong>estalia Fennica<br />

235, 68 p.<br />

PERTTUNEN, J., SIEVANEN, R., SALMINEN, H., VAKEVA, J., NIKINMAA, E., and SAARENMAA, H. (1994). i<br />

generic growth model f<strong>or</strong> a tree based on simple structural units and object-<strong>or</strong>iented programming. 30 p. (Subn-_tted).<br />

PRUNSINKIEWICZ, E and HANAN, J. 1989. Lindenmayer systems, fractals, and plants. Lecture Notes in Biomathernat_<br />

its 79, 120 p.<br />

de REFFYE, P., LECOUSTRE, R., EDELIN, C., and DINOUARD, P. 1989. Modelling plant growth and architecture, p.<br />

237-246. In Cell to Cell Signalling: from Experiments to The<strong>or</strong>etical Models. Academic Press, Oxf<strong>or</strong>d.<br />

66


RUMBAUGH, J., BLAHA, M., PREMERLANI, W., EDDY, E. and LORENSEN, W. 1991. Object-<strong>or</strong>iented modeling and<br />

design. 50() p. Prentice-Hall, Engtewood Cliffs, New Jersey.<br />

SAARENMAA, H. 1991. Object-<strong>or</strong>iented design of topological tree and f<strong>or</strong>est models, 2 p. In Proceedings of the First<br />

European Symposium on Terrestrial Ecosystems: F<strong>or</strong>ests and Woodlands. Fl<strong>or</strong>ence, Italy, May 20-24, 1991.<br />

SAA_NMAA, H., PERTTUNEN, J., VAKEVA, J., and NIKULA, A. 1994. Object-<strong>or</strong>iented modeling of the tasks and<br />

agents in integrated f<strong>or</strong>est health management. AI Applications 8(1): (In <strong>print</strong>).<br />

SALMINEN, H., SAARENMAA, H., PERTTUNEN, J., and VAKEVA, J. 1994. Modelling the hierarchical structure of<br />

trees with an object-<strong>or</strong>iented scheme, 18 p. In Resource Technology '94. International Symposium in Natural<br />

Resource Management (In <strong>print</strong>).<br />

SEQUEIRA, R.A., SHARPE, RJ.H., STONE, N.D., EL-ZlK, K.M., and MAKELA, M.E. 1991. Object-<strong>or</strong>iented simulation:<br />

plant growth and discrete <strong>or</strong>gan to <strong>or</strong>gan interactions. Ecol. Model. 58: 55-89.<br />

SIEVANEN, R. 1992. Construction and identification of models f<strong>or</strong> tree and stand growth. Helsinki university of technology.<br />

Automation _1_chnology Lab<strong>or</strong>at<strong>or</strong>y. Series A: <strong>Research</strong> Rep<strong>or</strong>ts 9, 155 p.<br />

TORASSO, R and CONSOLE, L. 1989. Diagnostic problem solving. 243 p. N<strong>or</strong>th Oxf<strong>or</strong>d Academic, London.<br />

TROENG, E. and LANGSTROM, B. 1991. Gas exchange in young Scots pine following pruning of current shoots. Ann.<br />

Sci. F<strong>or</strong>n. (Paris) 48: 359-366.<br />

67


DEFOLIATION OF FAGUS CRENATA AFFECTS THE POPULATION<br />

DYNAMICS OF THE BEECH CATERPILLAR,<br />

QUADRICALCARIFERA PUNCTATELLA<br />

NAOTO KAMATA t, YUTAKA IGARASHI I, and SEIJI OHARA 2<br />

_Lab<strong>or</strong>at<strong>or</strong>y of F<strong>or</strong>est Entomology, Tohoku <strong>Research</strong> Center<br />

F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute<br />

Nabeyashiki 72, Shimokuriyagawa, M<strong>or</strong>ioka, Iwate 020-01, Japan<br />

2Lab<strong>or</strong>at<strong>or</strong>y of Chemical Conversion, Wood Chemistry Division<br />

F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute<br />

Matshunosato 1, Kukizaki, Inashiki, Ibaraki 305, Japan<br />

INTRODUCTION<br />

In spite of remarkable progress in understanding the ecology of plant-herbiv<strong>or</strong>e interactions during the past 90 years,<br />

there is little yet known about the interactions between beeches and their foliv<strong>or</strong>es (Edwards and Wratten 1983, Edwards et<br />

al. 1986). Two beeches, Fagus crenata and E japonica, are dominant in the deciduous, broad-leaved f<strong>or</strong>ests of the cooltemperature<br />

zone in Japan. They are found from the southern part of Hokkaido, the n<strong>or</strong>thernmost main island, to the high<br />

mountains of Kyushu, the southernmost. The beech caterpillar, Quadricalcarifera punctatella (Motschulsky) (Lepidoptera:<br />

Notodontidae), is a monophagous species which feeds on beech leaves and occasionally completely defoliates the trees. V_e<br />

studied its population dynamics and tested the maternal effects hypothesis (Rossiter 1992) and the inducible resistance<br />

hypothesis, assuming that food deteri<strong>or</strong>ation occurs in beech following herbiv<strong>or</strong>y. A further objective was to determine the<br />

exact mechanisms by which Q. punctatella body size and density decrease after an outbreak has occurred.<br />

Life Hist<strong>or</strong>y and Features of Outbreaks<br />

Q. punctatella is a univoltine species (Igarashi, 1975). The pupa overwinters in the ground, and the adult emerges<br />

from late May to late July with a peak in mid- to late June (Kamata and Igarashi 1995a). Eggs are laid in masses on the<br />

underside of beech leaves, the total number per female being about 350. These are laid in several masses, each containing<br />

20-100 eggs (Kamata and Igarashi 1995b). The larva feeds only on beech leaves and molts 3 <strong>or</strong> 4 times. The peak appearance<br />

of the last instar larva is in early August.<br />

Serious outbreaks have been rec<strong>or</strong>ded since 1917, and outbreaks are known from central Honshu, the main island of<br />

Japan, to southern Hokkaido (Fig. 1) (Kamata and Igarashi 1995d). Most intervals between outbreaks have been 8-12 years<br />

and the average duration 1-4 years. Outbreaks in many regions have been synchronized even though the period of one <strong>or</strong> the<br />

time between two differed by time and place.<br />

METHODS<br />

Study Sites<br />

Larval density was estimated in Hakkohda (Site A), Hachimantai (Site B), and Appi (Site C), located in the n<strong>or</strong>thern<br />

part of Honshu island (Fig. 1). Outbreaks have been rec<strong>or</strong>ded at Sites A and B, but no conspicuous defoliation has ever been<br />

rec<strong>or</strong>ded in Site C (Kamata et al. in preparation).<br />

Mattson, W.J., Niemelii, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. U_IS)_A<br />

F<strong>or</strong>. Serv. Gen. Tech. Rep, NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

68


oo0<br />

a b<br />

ol<br />

¢' _achiman_i ()® 1,eAppi (C_<br />

Figure 1.---Locations of beech f<strong>or</strong>ests which have had been conspicuously defoliated by Q. punctatelIa (dark shading), and<br />

three study sites: Hakkohda (Site A), Hachimantai (Site B), and Appi (Site C).<br />

Estimation of Larval Density<br />

The density of Q. punctatella in each plot was estimated by trapping frass pellets of the last instar larvae, because<br />

defoliation is mostly the result of its feeding (Kamata and Yanbe, 1993). The density can usually be estimated by the<br />

following equation: Density = Total number of frass pellets per square meter/630, where 630 is mean number of frass<br />

pellets during last instar stage. Because density is underestimated when larval density becomes high and conspicuous<br />

defbliation occurs, the density must be calculated by the following equation: log,0 (Density) = 0.988"1og,0 X + 0.898,<br />

where X is the maximum number of frass pellets per day per square meter (Kamata and Igarashi, 1994a).<br />

Adult Number and Size<br />

Adult numbers were measured at Site B. Two light traps were set in an open area in a small valley surrounded by a<br />

f<strong>or</strong>est of predominantly beech. The distance between the two traps was 10m and both were 20m from the nearest edge of<br />

the beech f<strong>or</strong>est where the heaviest defoliation had occurred in a 1981 outbreak. Each light trap had two blue flu<strong>or</strong>escent<br />

bulbs (20W) and a basin with water and kerosene. The relative numbers were estimated by the total catches during the<br />

20 days from June 1lth to 30th, because the peak of the adult catches was in mid- to late June (Kamata and Igarashi,<br />

1994). Body size of the adult greatly influenced fecundity, but the fecundity of a trapped female cannot be estimated<br />

directly <strong>or</strong> indirectly from body weight, because both measurements decrease rapidly after emergence due to oviposition<br />

(Kamata and Igarashi 1995a). Rather, the area of f<strong>or</strong>ewing, which showed a good c<strong>or</strong>relation with both fecundity and<br />

adult weight, was used as a substitute measurement of body size and fecundity (Kamata and Igarashi 1995c). The area<br />

was theref<strong>or</strong>e measured by digitizer to the nearest mm z and its annual change was compared with population dynamics.<br />

In Site A, one light trap, which had two blue flu<strong>or</strong>escent bulbs (20W) and no basin, was set in a bare area ca. 20 m apart<br />

from the edge of beech f<strong>or</strong>est completely defoliated in 1990. A sheet of white, finely-meshed cotton cloth (1.8m x 1.8m)<br />

was set behind the trap and adults were caught on it. This survey was conducted several days each year from 1988 to<br />

1993. Area of the beech caterpillar's f<strong>or</strong>ewing was also measured. 69<br />

\


Defoliation Experiments<br />

To test the food deteri<strong>or</strong>ation hypothesis, the growth of larvae reared on beech saplings which had suffered previous<br />

artificial defoliation wascompared with controls. It was assumed that the effects of defoliation would be apparent in mid<br />

August. Treatments were a 2 year regime where all leaves were clipped f<strong>or</strong> 2 years running in 1991 and 1992. Another<br />

regime was a single year treatment during which all leaves were clipped f<strong>or</strong> only the year (1992) pri<strong>or</strong> to the experiment. No<br />

clipping was done in thecontrol group. Q. punctatella were reared in screened enclosures on branches of treated and control<br />

beeches. Larvae were occasionally shifted from one branch to another so that food was always plentiful. Surviv<strong>or</strong>ship and<br />

mature body size were compared by rearing the larvae to maturity on these saplings. The Q. punctatella used in the experiment<br />

<strong>or</strong>iginated from a population on Site A three generations after the previous outbreak.<br />

Testsf<strong>or</strong> Food Deteri<strong>or</strong>ation and Maternal Effects Hypotheses in the Field<br />

Three different regional Q. punctatella populations were reared on beech trees in the Tohoku <strong>Research</strong> Center of the<br />

F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute in M<strong>or</strong>ioka to evaluate population quality (Fig. 1). Larvae were reared in<br />

screened enclosures on branches and mature body sizes were compared.<br />

Eggs collected from these three regions were attached to beech saplings in each region. Larvae were also reared in<br />

screened enclosures on branches, and surviv<strong>or</strong>ship and mature body size compared among the three. From these results the<br />

influence of I_ostand herbiv<strong>or</strong>e quality on population dynamics and associated changes in body size were determined.<br />

Quantitative and Qualitative Properties of Beech Leaves<br />

_Ik)learn about changes in leaves which had been defoliated, their number and dry weight were determined. Dry<br />

weight of l0 leaf disks (2 cm in diameter), one punched from each of 10 randomly selected leaves, was viewed as an index of<br />

leaf density. Decrease inthe weight and the density of a leaf were compared to determine how leaves change due to defoliation.<br />

Leaf toughness wasalso measured using a "penetrometer" as described by Feeny (1970).<br />

We used the protein-precipitation assay as a measurement of the tannins. We measured the precipitation of BSA by<br />

extracts of beech leaves acc<strong>or</strong>ding to Martin and Martin (1983) and Makkar et al. (1987). Leaf powder (150 mg) was<br />

extracted three times with 3 ml of 70% acetone, and the total volume of the combined solution of extracts was adjusted to 10<br />

ml with the solvent. One ml of the above stock solution from leaf powder was added to 2 ml <strong>or</strong> 3 ml of 0.1% BSA solution<br />

(containing 2 mg <strong>or</strong> 3 mg of BSA, respectively) in acetate buffer (pH=5.0). The amount of precipitated BSA was determined<br />

by ninhydrin assay.<br />

RESULTS<br />

Population Dynamics and Body Size<br />

Larval densities changed with the same pattern in all three areas (Fig. 2). Density was lowest in 1986, and increased<br />

successively until 1990 (Kamata and Igarashi 1995c). In Site A, slight defoliation was recognized in 1989 and heavy<br />

defoliation in 1990. The predat<strong>or</strong>, Calosoma maximowiczi (Coleoptera, Carabidae), and an entomopathogenic fungus,<br />

C<strong>or</strong>dyceps militatis (Clavicipitales: Clavicipitaceae), greatly increased in abundance during the outbreak periods (Kamata<br />

and Igarashi 1994c, Sato et al. 1994). The densities in Sites B and C were not as high in 1990, and neither defoliation n<strong>or</strong><br />

natural enemies were conspicuous. However, the density decreased in 1991 in all three areas, then increased again in 1992.<br />

The resilience of the population density in 1992 was stronger in Sites B and C, where natural enemies were not conspicuous<br />

in 1990, than in Site A.<br />

Moth numbers in Site B fluctuated in tandem with larval density (Fig. 3). They were lowest in 1987, then increased.<br />

Interestingly, moth numbers were almost the same in 1990 and 1991, though the larval density decreased in 1991. Adult<br />

number began to decrease after 1992.<br />

70


1000 Conspicuous Defo i at ion<br />

100 .__. _; Si t e-A<br />

_.i Hakkohda<br />

t0 _ .... ,, _ i<br />

0.1<br />

o.o_i<br />

100 , i<br />

i : , Site-C<br />

10 _ _" Appi<br />

i ''_ " i....*<br />

>" i...*"! i<br />

,J<br />

o3<br />

0.1<br />

o.ol,iii i i ', i<br />

I00 i<br />

0.1 Site-B<br />

Hachimantai<br />

O.Ol i<br />

i<br />

o.ool ,,.... i ..... i i _ i<br />

198586 87 88 89 90 91 92 93<br />

Year<br />

Figure 2._Annual changes in last instar larval densities of Q. punctatella in the three study plots (mean + 90% confidence<br />

intervals). Site A (Hakkohda) is where conspicuous defoliation occurred during our study, Site B (Hachimantai) is<br />

where conspicuous defoliation had been rec<strong>or</strong>ded but did not occur during our study, and Site C (Appi) is where no<br />

conspicuous defoliation of <strong>this</strong> species has been rec<strong>or</strong>ded.<br />

71


400 -0<br />

" 350<br />

r'''"<br />

= 300<br />

.¢Z:<br />

,_ 250<br />

"='200<br />

E<br />

Femal e • :<br />

•<br />

'. , •<br />

"- 150<br />

o 100<br />

,5<br />

z 50 _.--_<br />

1600<br />

1400<br />

% 1200<br />

< 1000<br />

_._<br />

% 800<br />

Male<br />

! !<br />

_..===J<br />

." "_<br />

'-,--o600 _" "", ,<br />

_; 400 -/<br />

z 200<br />

' I _I I II il ! _I<br />

1985 1986 1987 1988 1989 1990 1991 1992 1993<br />

Year<br />

Figure 3.--Number of Q. punctatella adults caught in two light traps in Site B (Hachimantai) 1985-1993. The number of<br />

adults of each sex in each trap is shown. Nitrogen content of foliage was measured by a CHNS/O Analyzer<br />

(Perkin-Ehner PE2400 Series II) to the nearest 0.01%.<br />

Body size of moths gradually increased with larval density, but decreased suddenly in 1991 in Site A where severe<br />

defoliation had occurred the previous year (Fig. 4). A quantitative food sh<strong>or</strong>tage in 1990 was probably the main fact<strong>or</strong><br />

causing <strong>this</strong> size reduction. Body size increased a bit but was still small in 1992 and 1993, but larval density decreased and<br />

no food sh<strong>or</strong>tage occurred after 1991. Thus, fact<strong>or</strong>s other than a sh<strong>or</strong>tage of food were involved in the size decrease. Because<br />

size was ahnost the same in the two sites bef<strong>or</strong>e the outbreak, the great difference between the two after the outbreak<br />

was not genetically determined.<br />

72


2.oo 2.oo<br />

o t 4 t<br />

¢._ e'-<br />

1.80 1.80 ! t ! {<br />

_ _-<br />

,l--<br />

3 3:<br />

,-,- o° 1.60 Site-A t ,,° o 1, 60 Site-B Female<br />

I. 40 Female "-- o 1. 40<br />

a,_ 0.><br />

< I. 20 ..................... --'--- ........ < I. 20 ...............<br />

.--- 1 60 _ 1 60<br />

"E "E<br />

"; 140 "; 1 40 | I<br />

I<br />

g<br />

L I L_<br />

o<br />

4--<br />

1 30 Male t "<br />

4"- 0<br />

Male<br />

Site-A I ° 1 30 Site-B<br />

o 120 _ 1 20<br />

CL_<br />

110 ...... _--- ' ' < I 10 ----'---' ' ' ' ' " " "<br />

89 90 91 92 93 85 B6 87 88 89 90 91 92 93<br />

Year Year<br />

Figure 4.--Annual changes in adult size of (2. punctatella caught in light traps in Site A (Hakkohda) and Site B (Hachimantai)<br />

(mean + 95% confidence intervals). Area of f<strong>or</strong>e wing, which has a strong relationship to both fecundity and adult<br />

weight, is shown f<strong>or</strong> each sex.<br />

Defoliation Experiments<br />

Quantitative Changes in Beech Leaves and Food Limitation During an Outbreak Period<br />

The mean size and biomass of leaves declined precipitously after defoliation (Fig. 5). The 1-year treatments caused a<br />

46.1% decline (ca. 33.4 to 18.0 rag) in mean leaf dry weight from 1991-1992, and a 32.3% decline (ca. 41.2 to 30.8 mg)<br />

from 1992-1993. Following the second year of defoliation (i.e., in 1992), biomass declined further to ca. 12 rag/leaf in 1993,<br />

a 64.9% drop from the 1991 mean weight (Fig. 5a). Leaf density <strong>or</strong> dry weight of 10 leaf disks (31.4 cm 2total) from defoli ....<br />

ated trees followed the same pattern: 1-year treatments caused a 14.0% decline (108 to 92.8 rag/10 disks) from 1991-1992,<br />

and an 8.0% decline (104 to 93.8 rag/10 disks) from 1992-1993 (Fig. 5b). The 2-year treatment caused a further decline in<br />

leaf density to ca. 81 rag/10 disks, > 25% down from the initial mass in 1991 (Fig. 5b). Besides a decline in leaf size, and<br />

leaf density, the number of leaves on each tree also decreased slightly after the 1-year treatment, though two of 12 trees<br />

exhibited a slight increase in leaves (Fig. 5c). Consequently, total leaf biomass per tree declined by about 50% due to the 1year<br />

treatment (Fig. 5d).<br />

In two plots at Site A, slight defoliation was observed in 1989, the amount of leaves fed on by Q. punctatella was<br />

estimated at ca. 60% of total canopy, and ca. 20% of the beech trees were completely defoliated. The following year, Q.<br />

punctatella increased in density and defoliated completely ahnost all beech trees. Food limitation was an imp<strong>or</strong>tant m<strong>or</strong>tality<br />

fact<strong>or</strong> in <strong>this</strong> generation because many dead larvae, which had not been killed by predat<strong>or</strong>s <strong>or</strong> pathogens and were supposed<br />

to have died of starvation, were scattered on the ground. Considering the results of the defoliation experiment, the biomass of<br />

leaves should have decreased in 1990 as a result of the severe defoliation the previous year. A negative feedback occurred:<br />

as herbiv<strong>or</strong>e density gradually increased the defoliation reduced the amount of available food in the following generation,<br />

thus acting as a limiting fact<strong>or</strong>.<br />

73


co 100-- d<br />

><br />

_' 10<br />

01--I<br />

700 --<br />

I I<br />

60 CD 600--<br />

500-<br />

400c<br />

_-- o 300 -<br />

200z<br />

I00- - -"<br />

60<br />

60<br />

._<br />

0--1<br />

120--<br />

1 I<br />

110<br />

•--- _ 100 @ Control<br />

_- E _[_ 1 Yr<br />

o "--- 90 -- Treatment<br />

2Yr<br />

"=: 80 -- Treatment<br />

(3,}<br />

_J<br />

70-I I I I<br />

50 Control<br />

o. __ _E 40<br />

30<br />

L<br />

l -""_ 1 Yr<br />

.= 20--<br />

.,.., "---" l _ i l Treatment<br />

' _ 2 Yr<br />

"_ 10_ _ Treatment<br />

0---I I I I<br />

1991 1992 1993<br />

Year<br />

Figure 5.--Quantitative changes in beech leaves in the year following artificial defoliation, a: Dry weight per leaf, b: Index.<br />

74<br />

of leaf density measured by dry weight of 10 leaf disks (2 cm in diameter), c: Number of leaves in each tree, d: Dry<br />

weight of all leaves in each tree.


Qualitative Changes in [,eaves and Larval Growth<br />

Leaf"toughness, like leaf"density, decreased following clipping of the leaves: it was only 85.7% of the control the<br />

year following the 1 year treatment and 69.8% of the control after the 2 year treatment (Fig. 6). Nitrogen content was almost<br />

the same in untreated trees tot 3 years (1991-1993) but decreased after leaves were clipped. It was 2.23-2.26% bef<strong>or</strong>e the<br />

treatment, 2.00-2.05% the year following the 1 year treatment (ca. 10% reduction), and 1.88% the year following the 2 year<br />

treatment (16.6% reduction) (Fig. 7a). The N content following the 1 year treatment was almost the same in 1992 and 1993.<br />

These results clearly indicate that nitrogen content is reduced by defoliation and is influenced not by the year but by the<br />

degree of defoliation. Tannin content, however, increased the year following leaf clipping (Fig. 7b). When 2 mg of BSA was<br />

added, about 0.4 mg of the protein was precipitated in the regimes of 2 year treatment and 1 year treatment, although precipitation<br />

was not detected in the control regime. Furtherm<strong>or</strong>e, when 3 mg of BSA was added, no precipitation was recognized<br />

in the 1 year treatment <strong>or</strong> the control regimes, but 0.43 mg of the protein was precipitated following the 2 year treatment.<br />

These quantitative and qualitative changes in beech following defoliation supp<strong>or</strong>t that the trees were less vig<strong>or</strong>ous after<br />

defoliation, and also less suitable f<strong>or</strong> herbiv<strong>or</strong>es because of the declining nitrogen and increasing tannins.<br />

180<br />

160 I "<br />

z J 153.5<br />

[<br />

co |<br />

140<br />

131.7<br />

c<br />

= 120<br />

0<br />

F---<br />

100<br />

(:D<br />

107.2<br />

L_ ............. 1<br />

--J T<br />

80 ' ' ................<br />

2 Yr Treatment 1 Yr Treatment C0ntr0]<br />

Figure 6._Leaf toughness of beechleavesin the year following artificial defoliation. The number is a mean value, bar<br />

means maximum and minimum value, and upper and lower sides of box are one standard err<strong>or</strong>.<br />

Surviv<strong>or</strong>ship rates on beech saplings were determined using the number of hatched larvae as a base. The x-axis in<br />

Figure 8 is the number of days following egg attachment on June 30th. Although approximately half of the larvae reared on<br />

the control saplings survived to maturity, only about 10% of those reared on the 2 year treatment trees survived to maturity.<br />

Many of the latter died during early stages of larval development. Surviv<strong>or</strong>ship on the 1 year treatment was intermediate<br />

between the 2 year treatment and the control. Both female and male body size declined f<strong>or</strong> larvae reared on clipped saplings,<br />

as did adult body size (Fig. 9). In the 2 year treatment regime, body size was smaller than of the 1 year treatment regime, and<br />

was just three quarters of that in the control. Defoliation lowers beech quality which, in turn, lowers larval surviv<strong>or</strong>ship and<br />

body size.<br />

Population Quality of Herbiv<strong>or</strong>e in Three Sites<br />

Test f<strong>or</strong> Maternal Effects and Food Deteri<strong>or</strong>ation in the Field<br />

The larvae from Site A had smaller body size (p < 0.01 t-test) than those from the other two regions (Fig. 10). This<br />

tendency was exactly the same f<strong>or</strong> males and females, as well as f<strong>or</strong> mature larvae and f<strong>or</strong> pupae. No difference was found<br />

between the Site B and C populations. Thus it can be said that even three generations after an outbreak, the quality of the<br />

Site A population was less than that of the populations from the other two regions where density was increasing.<br />

75


0.5 b<br />

O.4 AddedBSA<br />

2mg<br />

=0.3<br />

,_ El 3mg<br />

-'0.2<br />

- r--"<br />

o0.1<br />

'- N.D. N.D<br />

_ j = 1<br />

2 Yr Tr 1 Yr Tr Control<br />

2.3-- a<br />

_= _<br />

o<br />

2.2--<br />

2.1-<br />

{ Control<br />

1Yr<br />

z<br />

1 9-<br />

"<br />

Treatment<br />

Yr<br />

Tre;._tment<br />

1"77 I I I<br />

1991 1992 1993<br />

Year<br />

Figure 7.--Nitrogen and tannins contained in beech leaves the year following artificial defoliation, a: Total nitrogen, b:<br />

Weight of precipitated BSA by addition of 2 ml of 0.1% BSA solution (c<strong>or</strong>responding to addition of 2 mg of<br />

BSA) and 3 ml of 0.1% BSA solution (3 mg of B SA). N.D. indicates that no precipitation was detected.


t00 2 Yr 1 Yr<br />

o_ 80 Treatment Treatment . :_,,"<br />

_- 60<br />

o 40<br />

o_ 20<br />

Control<br />

-- L_<br />

0 20 40 60 80 0 20 40 60 80 0 20 40 60 80<br />

Days from Jun. 30<br />

Figure 8.--Surviv<strong>or</strong>ship curves of Q. punctatella larvae reared on beech which had suffered artificial defoliation. Each line<br />

began with an egg-mass containing about 50 eggs. X-axis is the number of days following egg attachment on June<br />

30th.<br />

500 9,_<br />

400 I l<br />

--_ [----_ w<br />

300<br />

¢....<br />

200<br />

; 291.5 345 392.9 Mean<br />

100 2 8 34 N<br />

_ !<br />

450<br />

4O0<br />

o_<br />

|<br />

350 [ I<br />

.....'_300 F ---_L--- ..... ] I<br />

250<br />

•-'= 200<br />

l ' n l<br />

"T_150 236.7 289.5 353.3 Mean<br />

3= 100<br />

50<br />

3 10 38 N<br />

! _I _1<br />

2 Yr Treatment 1 Yr Treatment Control<br />

Figure 9.--Body size of Q. pttnctatella larvae reared on beech which had suffered artificial defoliation. Weight of larvae at<br />

maturation is indicated f<strong>or</strong> each sex. Bar means maximum and minimum value, and upper and lower sides of box are<br />

one standard err<strong>or</strong>. 77


__ Maturation 500 o_ Maturation<br />

_ 600 E<br />

r-- I _ --- 300<br />

400 -'=<br />

447.1 507.5 504.7 Mean .=,,_200 339 7 425 6 405.7 Mean<br />

• •<br />

200 30 21 19 N _= 100 37 27 20 N<br />

0<br />

_t Pupation o_ Pupation<br />

600 500<br />

500 ]....._ I --- 400<br />

40o I Ij F-__] , 3oo I 1 w<br />

300 El, _ 200 '<br />

._ 200 342.6 396.6 406.3 Mean 265.3 339.1 329.6 I_ean<br />

100 29 21 19 N _= 100 32 27 20 II J<br />

0 , , , 0 ' '<br />

A B C A B C<br />

Locationof Q. punctatella Locationof Q. punctatella<br />

Figure lO.---Body size of three different regional Q. punctatella populations reared on beeches in the Tohoku <strong>Research</strong><br />

Center of the F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute in M<strong>or</strong>ioka. Weight of larvae at maturation and at<br />

pupation are indicated f<strong>or</strong> each sex. Bar means maximum and minimum value, and upper and lower sides of box are<br />

one standard err<strong>or</strong>.<br />

Tests of Two Hypotheses in Field Populations<br />

Larvae from all of the regions were reared on beech saplings in each of the respective regions. Plant quality influenced<br />

surviv<strong>or</strong>ship in each region. Although great differences in surviv<strong>or</strong>ship between regions were observed, little difference<br />

was found within a region (Table 1). Approximately half of the larvae reared on beech saplings in Sites B and C<br />

survived to maturity, whereas surviv<strong>or</strong>ship of all those reared on Site A beech was between 3.2% and 6.5%. The influence of<br />

Q, p,nctawlla population quality in each region on body size was clearly evident. Irrespective of the location of beech<br />

saplings on which the larvae were reared, the <strong>or</strong>der of average body size among the regional populations was the same: B >__<br />

C 2 A (Table 2). In particular, body size of A population larvae was significantly smaller than that of larvae from the other<br />

Table !.--Surviv<strong>or</strong>ship of Q. punctatella larvae from each region on beech in that region.<br />

78<br />

Q. punctatella<br />

A B C<br />

A 3.6% 3.2% 6.5%<br />

(6/165) (5/155) (10/154)<br />

E crenata B 48.9% 55.0% 53.9%<br />

(44/90) (33/60) (41/76)<br />

C 55.8% 52.9%<br />

(72/129) (36/68)<br />

(No. of matured larvae / No. of hatched larvae)


Table 2.--Body size of Q. purzctatella from each region on beech in that region. Weight of larvae at maturation is shown f<strong>or</strong><br />

each sex.<br />

Female Q. purzctatella<br />

A B C<br />

376.3+68.5 442.0+ 15.6 414.4+48.9<br />

A 388 453 489<br />

369 431 360<br />

3 2 5<br />

398.5+75.3 484.5+108.9 451.3_+58.8<br />

E crenata B 576 739 554<br />

309 333 340<br />

17 17 22<br />

392.9+26.3 455.0+68.5<br />

c 451 550<br />

351 342<br />

34 14<br />

Male Q. punctatella<br />

A B C<br />

288.3+60.5 369.7+ 15.6 336.8+41.3<br />

A 349 384 398<br />

228 353 298<br />

3 3 5<br />

294.5+63.6 384.9+36.99 335.3+45.2<br />

E crenata B 416 469 409<br />

204 332 254<br />

27 16 19<br />

353.3+31.9 396.0+-39.0<br />

C 412 450<br />

279 299<br />

38 22<br />

Mean+SD<br />

Max<br />

Min<br />

N<br />

two regional populations (p < 0.01 t-test). The influence of beech on body size was not as apparent as the population effect:<br />

female body size of larvae reared on Site B beech was close to that of larvae reared on Site C beech, while male body size of<br />

larvae reared on Site B beech was similar to that of those reared on Site A beech.<br />

Leaf properties were compared among the three sites (Fig. l 1), and weight per leaf among the three was almost the<br />

same. Leaves of Site A beech were significantly tougher than those of the other two sites, though they were much thinner<br />

than Site C leaves and almost the same as the Site B leaves. A remarkable point is that nitrogen content was significantly<br />

higher in leaves of Site A than in those of the other two sites (Fig. 12). Tannin content also increased greatly. Judging from<br />

the high N content of beeches in Site A, where severe defoliation had occurred three years bef<strong>or</strong>e our experiment, trees had<br />

79


240<br />

__ 220<br />

O_<br />

200<br />

180<br />

C<br />

= 0 160<br />

140<br />

4--<br />

" 120<br />

.J<br />

100 ......<br />

..<br />

153.5<br />

.110 b<br />

__,_<br />

_100 _____[ w<br />

90 88.3 89.7 101.7<br />

70 j ! I<br />

100<br />

a<br />

4-- 80 I<br />

,_J<br />

46.5<br />

,=- 4O 42.3<br />

-,-" °o,o 1 ....... • I 49 2<br />

.'=" 20 7-----<br />

.............. 1 i a<br />

A B C<br />

Site<br />

Figure l l._Leaf thickness and toughness of beech leaves in three different regions. Leaf thickness was represented by dry<br />

weight of 10 leaf disks (2 cm in diameter). The number is a mean value, bar means maximum and minimum value,<br />

and upper and lower sides of box are one standard err<strong>or</strong>.<br />

80


06 b<br />

AddedBSA<br />

< 04 _ 2rag<br />

-'=' i:3<br />

-,03<br />

'F=,.0 2<br />

_-01<br />

_ N.D.<br />

,_ _ _ J<br />

3 a<br />

2.8 ...... _<br />

2.84<br />

_2.6 i .............<br />

=" 2.4 2.58<br />

...... ..... i------=-_- _<br />

2.2 ] ! -- 2.36 w I iiiiii<br />

....... ,..... L_.__J"....... I _ _ H _, I _'_ " _fl<br />

A B C<br />

Sit,e<br />

Figure 12.--Nitrogen ahd tannins contained in beech leaves in three different regions, a: Total nitrogen (The number is a<br />

mean value, bar means maximum and minimum value, and upper and lower sides of box are one standard err<strong>or</strong>), b:<br />

Weight of precipitated BSA when added at 2 mg and 3 mg BSA. N.D. indicates that no precipitation was detected.<br />

become vig<strong>or</strong>ous. However, since the trees had also increased their defensive secondary compounds, they had also become<br />

less suitable f<strong>or</strong> herbiv<strong>or</strong>es, and insect m<strong>or</strong>tality was higher than in trees in the other two sites.<br />

DISCUSSION<br />

Self regulation by reducing adult body size and curtailing egg production in insects during an outbreak with resultant<br />

declines in the initial number of individuals in the next generation is thought to be one of the maj<strong>or</strong> fact<strong>or</strong>s governing the<br />

population dynamics of insects (Klomp 1966, Dempster and Pollard 1981). This has been shown to be a consequence,<br />

among other fact<strong>or</strong>s, of insufficient food supply due to complete defoliation of the host, as well as a density dependent<br />

outcome resulting from conditions experienced during larval development. However, <strong>this</strong> phenomenon is not restricted<br />

81


solely to the outbreak generation. Many insects have been shown to have body size changes associated with their population<br />

density dynamics. In other w<strong>or</strong>ds, during a period when population density is increasing, body size also increases, and the<br />

reverse occurs when population density is declining, body size decreases also (Baltensweiler and Fischlin 1988, Myers 1990),<br />

In Q. punctatella, body size changed in the same way as the density. Changes in fecundity caused by these changes in body<br />

size were probably one fact<strong>or</strong> causing the population fluctuation.<br />

Because the body size of Q. punctatella at Site A bef<strong>or</strong>e the outbreak was approximately the same as that f<strong>or</strong> the Site<br />

B population, the current small body size of the f<strong>or</strong>mer is thought not to be genetically determined. Body size at Site A<br />

dropped to 70% of the n<strong>or</strong>m the year following the outbreak and was 80% even two generations after the outbreak. This<br />

cannot not be explained solely by a lack of food during the outbreak. The results indicate that beech which has experienced<br />

de_i)liation the previous year is a po<strong>or</strong> host, causing both surviv<strong>or</strong>ship and body size of larvae to decline. Even though 3<br />

years had passed since the outbreak at Site A, larvae reared on Site A beech still had low surviv<strong>or</strong>ship. C<strong>or</strong>responding to host<br />

quality, Q. punctatella quality was still bad in Site A. Even when Q. punctatella populations from <strong>this</strong> site were reared on<br />

beech from the other sites where density was increasing, the bodies of these larvae were still significantly smaller. High<br />

m<strong>or</strong>tality, due to deteri<strong>or</strong>ation of food and small body size, and low egg production due to po<strong>or</strong> Q. punctatella quality,<br />

explains why its density continued to decline years after the outbreak. Several auth<strong>or</strong>s have demonstrated that po<strong>or</strong> nutritional<br />

status in one generation may deter n<strong>or</strong>mal development in the next. Examples include Lymantria dispar (Kovasevic<br />

1956), Malacosoma pluviale (Wellington 1965), and Hyphantria cunea (M<strong>or</strong>ris 1967). Such intergenerational, cumulative<br />

effects are called "maternal effects" (Rossiter 1992). Rossiter (1992) states "maternal effects are the result of resource<br />

provisioning by one generation f<strong>or</strong> the next... The resource based maternal effects are the product of gene-environment<br />

interactions experienced in the parental generation." Deteri<strong>or</strong>ation in Q. punctatella quality after the outbreak was the very<br />

result caused by maternal effects.<br />

Haukiojia and Neuvonen (1987) proved the hypothesis that po<strong>or</strong> food may select f<strong>or</strong> individuals with superi<strong>or</strong> ability<br />

to process low-quality diets. However, experiments conducted here did not confirm <strong>this</strong>: the surviv<strong>or</strong>ship of Site-A Q.<br />

punctatetla was significantly lower than that of Site C Q. punctate[la, and was almost the same as that of Site B Q.<br />

punctatella on po<strong>or</strong>-quality Site A beech.<br />

Concerning the food deteri<strong>or</strong>ation hypothesis, it is necessary to distinguish between induced changes in secondary<br />

compounds such as tannin, which are an active means of defense (induced defense hypothesis) and that of passive changes in<br />

secondary chemistry following episodes of defoliation (Myers 1988). The defense strategy of beech, in relation to severe<br />

defoliation, changed as time passed after an outbreak, though in each case it resulted in high m<strong>or</strong>tality and small body size of<br />

Q. punctatella. In the year following defoliation, beech became less vig<strong>or</strong>ous and the defense response was a passive one<br />

(food deteri<strong>or</strong>ation): tannins increased and nitrogen content decreased. Next, trees became m<strong>or</strong>e vig<strong>or</strong>ous and the defensive<br />

strategy changed to a m<strong>or</strong>e active one; both nitrogen and tannin content increased.<br />

The results on Site B beech were unexplainable by <strong>this</strong> scenario; nitrogen content was lower than Site A beech and<br />

tannin content was almost the same, but the tree perf<strong>or</strong>mance was comparable to Site C beech f<strong>or</strong> rearing Q. punctatella.<br />

These results indicate that it is very dangerous to discuss the perf<strong>or</strong>mance of plants by measuring only nitrogen <strong>or</strong> secondary<br />

compounds. Clancy (!99 l) also demonstrated that Douglas fir trees susceptible to the western spruce budw<strong>or</strong>m had lower<br />

levels of foliar nitrogen and sugars than resistant trees, and that the susceptible trees had mineral/nitrogen ratios which were<br />

closer to optimal levels. Not only secondary compounds such as phenols and tannins but also minerals thus must be taken<br />

into consideration.<br />

Nitrogen content in untreated trees at Site A was nearly the same in all 3 years (1991-1993) and that in beeches<br />

following 1year of defoliation was the same f<strong>or</strong> 2 years (1992-1993). It differed greatly among the three sites, but variance<br />

within a site was very minimal. Beech leaves at outbreak Sites A and B showed higher nitrogen content than at the nonoutbreak<br />

Site C. There is a possibility that nitrogen level is linked to the site dependent outbreak characteristics of Q.<br />

punctatella.<br />

The induced defense hypothesis advocates the following scenario. Due to the decline in food quality after an<br />

outbreak, succeeding generations decline, that is, insect growth is checked and population numbers decrease. With the<br />

recovery of plant quality, insect population density begins to increase again, thus creating the cyclical population dynmnic:s<br />

of these insects. Two different types of beech response were recognized in the decline in food quality as time passed afte:r an<br />

outbreak (Fig. 13). Beech becomes less vig<strong>or</strong>ous soon after severe defoliation, and nitrogen content decreases but deferlsive<br />

82


Severe Defol iat ion Site-A 1993<br />

_ _ Tree vigour<br />

_ -<br />

j<br />

V i1<br />

i<br />

I<br />

V I<br />

_ ,,<br />

NitrogenContent<br />

Deiensive Secondary Compounds<br />

Food Quail ty<br />

oar<br />

Passive Defense -_-------_Active Defense<br />

Figure 13.--Schema of defense response of beech after severe defoliation.<br />

secondary compounds increase. This rather passive food deteri<strong>or</strong>ation causes high m<strong>or</strong>tality and smaller body size in the Q.<br />

punctatella population. Thus, Q. punctatella density decreases and beech is released from severe herbiv<strong>or</strong>y pressure. Beech<br />

gradually becomes vig<strong>or</strong>ous again and its defense strategy changes to a m<strong>or</strong>e active one; defensive compounds such as<br />

tannins increase, though nitrogen content also recovers. Such beeches are not suitable f<strong>or</strong> herbiv<strong>or</strong>es, which results in high<br />

m<strong>or</strong>tality of Q. punctatella.<br />

83


SUMMARY<br />

Body size of Q. punctatella at first increased with population density, but became very small the year following an<br />

outbreak. Its size was still small 3 years afterwards. Because adult size greatly influences fecundity, <strong>this</strong> was a fact<strong>or</strong> in<br />

keeping Q. punctatella density low f<strong>or</strong> several years after an outbreak. As f<strong>or</strong> food quantity, severe defoliation caused a<br />

decrease in the amount of leaves the following year, and <strong>this</strong> limited the peak density. The defense strategy of beech in<br />

relation to severe defoliation changed as time passed. In the year following severe defoliation, nitrogen content decreased<br />

and tannins increased; <strong>this</strong> resulted in high m<strong>or</strong>tality and small body size in Q. punctatella. Three years after an outbreak,<br />

nitrogen content recovered to the same level as pri<strong>or</strong> to defoliation, but tannins were even higher. This active, induced<br />

defense kept larval surviv<strong>or</strong>ship low. Maternal effects were also recognized; the quality of Q. punctatella was po<strong>or</strong> even 3<br />

years after an outbreak. Thus, food deteri<strong>or</strong>ation and the c<strong>or</strong>related results of maternal effects suppressed Q. punctatella<br />

density f<strong>or</strong> several years after an outbreak.<br />

ACKNOWLEDGMENTS<br />

We would like to sincerely thank Dr. Garry J. Piller (School of Environmental Earth Sciences, Hokkaido University)<br />

f<strong>or</strong> his kind advice and discussion.<br />

LITERATURE CITED<br />

BALTENSWEILER, W. and FISCHLIN, A. 1988. The larch bud moth in the Alps, p. 332-351. In Berryman A.A., ed.<br />

Dynamics of F<strong>or</strong>est Insect Populations. Plenum, New Y<strong>or</strong>k-London.<br />

CLANCY, K.M. 1991. Douglas-fir nutrients and terpenes as potential fact<strong>or</strong>s influencing western spruce budw<strong>or</strong>m defoliation,<br />

p. 124-134. In Baranchikov, Y.N., Mattson, W.J., Hain, F., and Payne, T.L., eds. F<strong>or</strong>est insect guilds: patterns of<br />

interaction with host trees. GTR-NE-153. Radn<strong>or</strong>, PA: U.S. Department of Agriculture, F<strong>or</strong>est Service.<br />

DEMPSTER, J.P. and POLLARD, E. 1981. Fluctuations in resource availability and insect populations. Oecologia 50: 412-<br />

416.<br />

EDWARDS, P.J. and WRATTEN, S.D. 1983. Wound induced defenses in plants and their consequences f<strong>or</strong> patterns of<br />

insect grazing. Oecologia 59: 88-93.<br />

EDWARDS, P.J., WRATTEN, S.D., and GREENWOOD, S. 1986. Palatability of British trees to insects: constitutive and<br />

induced defenses. Oecologia 69:316-319.<br />

FEENY, P. 1970. Seasonal changes in oak leaf tannins and nutrients as a cause of spring feeding by winter moth caterpillars.<br />

Ecology 51:565-581.<br />

HAUKIOJIA, E. and NEUVONEN, S. 1987. Insect population dynamics and induction of plant resistance: the testing of<br />

hypotheses, p.411-432. In Barbosa, E and Shultz, J.C., eds. Insect Outbreaks. Academic Press, New Y<strong>or</strong>k.<br />

IGARASHI, M. 1975. The beech caterpillar, Quadricalcarifera punctatella (Motschulsky)(Lep., Notodontidae), as an<br />

imp<strong>or</strong>tant defoliat<strong>or</strong> of beech, Fagus crenata Biume. Monthly Rep<strong>or</strong>t of Tohoku <strong>Research</strong> Center, F<strong>or</strong>estry and<br />

F<strong>or</strong>est Products <strong>Research</strong> Institute 162: I-4. l(in Japanese)<br />

KAMATA, N. and IGARASHI, Y. 1994. Problems in estimating the density of larval beech caterpillar, Quadricalcarifera<br />

punctatella (Motschulsky)(Lep., Notodontidae), using frass drops by modification of Southwood-Jepson's method. J.<br />

Appl. Ent. 118: 92-98.<br />

KAMATA, N. and IGARASHI, M. 1995a. Diurnal change of the adult behavi<strong>or</strong>, daily oviposition, and influences of<br />

temperature on adult emergence and light trapping data of the beech caterpillar, Quadricalcarifera punctatella<br />

(Motschulsky) (Lep: Notodontidae). J. Appl. Ent. 119: 177-184.<br />

84


KAMATA, N. and IGARASHI, M. 1995b. Relationship betweeen temperature, number of larval instars, larval growth, body<br />

size, and adult fecundity of the beech caterpillar, QuadricaIcarifera punctatella (Motschulsky) (Lep., Notodontidae):<br />

As a cost-benefit trade-off. Environ. Entomol. 24: 648-656.<br />

KAMATA, N. and IGARASHI, Y. 1995c. An example of numerical response of the carabid beetle, Calosoma maximowiczi<br />

M<strong>or</strong>awitz (Col., Carabidae), to the beech caterpillar, Quadricalcarifera punctatella (Motschulsky) (Lep.,<br />

Notodontidae). J. Appl. Ent. 119: 139-1142.<br />

KAMATA, N. and IGARASHI, Y. 1995d. Synchronous populations of the beech caterpillar, Quadricalcarifera punctatella<br />

(Motschulsky): Rainfall is the key, p. 452-473. In Hain, R.R., Salom, S., Ravlin, W., Payne, T., and Raffa, K., eds.<br />

Behavi<strong>or</strong>, population dynamics, and control of f<strong>or</strong>est insects. <strong>USDA</strong> F<strong>or</strong>est Service, Radn<strong>or</strong>, PA 19087.<br />

KAMATA, N. and YANBE, T. 1993. Frass production of last instar larvae of the beech caterpillar, Quadricalcarifera<br />

punctatella (Motschulsky)(Lep., Notodontidae), and method of estimating their density. J. Appl. Ent. 117:84-9 I.<br />

KLOMP, H. 1966. The dynamics of a field population of the pine looper (Lepi., Geom.). Adv. Ecol. Res. 3: 207-303.<br />

KOVASEVIC, Z. 1956. Die nahrungsswahl und das auftreten der pflanzenschadlinge. Anz. Schaedlingskd. 24: 97-101.<br />

MAKKAR, H.P.S., DAWRA, R.K., and SINGH, B. 1987. Protein precipitation assay f<strong>or</strong> quantitation of tannins: determination<br />

of protein in tannin-protein complex. Anal. Biochem. 166: 435-439.<br />

MARTIN, J.S. and MARTIN, M.M. 1983. Tannin assays in ecological studies. -Precipitation of Ribulose-l,5-Bisphosphate<br />

Carboxylase/Oxygenase by tannic acid, quebracho, and oak foliage extracts. J. Chem. Ecol. 9: 285-294.<br />

MORRIS, R.F. 1967. Influence of parental food quality on the survival of Hyphantria cunea. Can. Entomol. 99: 24-33.<br />

MYERS, J.H. 1988. The induced defense hypothesis: Does it apply to population dynamics of insects? p. 345-365. In<br />

Spencer, K.C., ed. Chemical Mediation of Coevolution. Academic Press, San Diego.<br />

MYERS, J.H. 1990. Population cycles of western tent caterpillars: experimental introductions and synchrony of fluctuations.<br />

Ecology 71" 986-995.<br />

ROSSITER, M.C. 1992. The impact of resource variation on population quality in herbiv<strong>or</strong>ous insects: a critical aspect of<br />

population dynamics, p. 13-42. In Hunter, M.D., Ohgushi, T., and Price, RW., eds. Effects of Resource Distribution<br />

on Animal-Plant Interactions. Academic Press, San Diego.<br />

SATO, H., SHIMAZU, M., and KAMATA, N. 1994. Detection of C<strong>or</strong>dyceps militalis Link (Clavicipitales: Clavicipitaceae)<br />

by burying pupae of Quadricalcarifera punctatella (Motschulsky) (Lepidoptera: Notodontidae). Appl. Ent. Zool. 29:<br />

130-132.<br />

WELLINGTON, W.G. 1965. Some maternal influences on progeny quality in the western tent caterpillar, Malacosoma<br />

pluviale (Dyar.). Can. Entomol. 97: 1-14.<br />

85


THE IMPACT OF FLOWERING ON THE SUITABILITY OF BALSAM FIR<br />

FOR SPRUCE BUDWORM VARIES WITH LARVAL FEEDING BEHAVIOR<br />

1_.BAUCE and N. CARISEY<br />

D6partement des Sciences F<strong>or</strong>esti_res, Facult6 de F<strong>or</strong>esterie et de G6omatique<br />

Universit6 Laval, Ste-Foy (Qu6bec), Canada G 1K 7P4<br />

INTRODUCTION<br />

The production of staminate flowers by coniferous trees is known to affect the fitness of insect defoliat<strong>or</strong>s. F<strong>or</strong><br />

instance, spruce budw<strong>or</strong>m, Ch<strong>or</strong>istoneurafumiferana Clem., larvae on flowering balsam fir, Abies balsamea (L.) Miller, had<br />

3 to 7 days sh<strong>or</strong>ter development times than those on nonflowering trees (Jaynes and Speers 1949, Blais 1952, Greenbank<br />

1963). Based on <strong>this</strong> effect, Blais (1952) hypothesized that flowering of balsam fir could trigger spruce budw<strong>or</strong>m outbreaks.<br />

The impact of flowering on budw<strong>or</strong>m survival and fecundity varies greatly from one study to another. Blais (1952)<br />

found no significant effect of flowering on spruce budw<strong>or</strong>m survival, while Mattson et al. (1991) detected a 25% increase in<br />

larval survival when insects were reared on balsam fir branches bearing staminate flowers compared to those on nonflowering<br />

branches. M<strong>or</strong>eover, balsam fir flowering had no effect on the fecundity of spruce budw<strong>or</strong>m (Blais 1952), whereas Pinus<br />

banksiana Lamb. flowering had a negative effect on the fecundity of the jack pine budw<strong>or</strong>m, Ch<strong>or</strong>istoneura pinus. Although<br />

there is no agreement on whether flowering positively <strong>or</strong> negatively affects budw<strong>or</strong>m m<strong>or</strong>tality and fecundity, most auth<strong>or</strong>s<br />

believe that the impact of flowering results from either the high nutritive value of the pollen and/<strong>or</strong> the presence of staminate<br />

flower clusters that may provide a m<strong>or</strong>e fav<strong>or</strong>able micro-habitat in term of heat and protection from natural enemies (Lejeune<br />

and Black 1950, Wellington 1950, Blais 1952, Greenbank 1963, Mattson etal. 1991).<br />

In regard to budw<strong>or</strong>m dynamics, the production of staminate flowers can be followed, the next year, by either an<br />

increase (Graham 1935, Blais 1952, Volney 1988) <strong>or</strong> a decrease (Batzer and Jennings 1980, Bauce 1986) in budw<strong>or</strong>m<br />

population density. In <strong>or</strong>der to explain such opposite trends, we hypothesize and test here that the impact of flowering on the<br />

suitability of balsam fir f<strong>or</strong> budw<strong>or</strong>m larvae varies acc<strong>or</strong>ding to the feeding behavi<strong>or</strong> of the larvae, which is affected by the<br />

density of budw<strong>or</strong>m population (Blais 1952).<br />

METHODS<br />

Field Rearing Experiment<br />

The impact of flowering on spruce budw<strong>or</strong>m growth, development, survival, and fecundity was determined by caging<br />

second instar larvae in the mid- and lower-crowns of flowering and nonflowering balsam fir trees. In 1992, a total of 10<br />

flowering and 10 nonflowering dominant balsam fir trees were randomly selected in a 60 yr-old balsam fir stand located in<br />

compartment 20 of the F<strong>or</strong>_t Montm<strong>or</strong>ency (47° 19' N, 79° 09' W), an experimental f<strong>or</strong>est of Universit6 Laval. The stand fits<br />

Grandtner's (1966) description of the balsam fir-white birch association and was classified as site 1 quality with good<br />

drainage (B61anger et al. 1983), 15%slope, deep uncompacted glacial till, and 65% crown cover.<br />

A severe drought in 1991 triggered an intense production of staminate flowers the following spring. Trees located<br />

near streams, <strong>or</strong> not fully exposed to sunlight did not produce staminate flowers. On each sample trees, two 90 cm long<br />

branches, facing n<strong>or</strong>th-n<strong>or</strong>thwest, were selected at the mid and lower third sections of the crown. Staminate flowers were<br />

present only in the midcrown of flowering trees. Reproductive buds burst betbre vegetative buds and the pollen was dispersed<br />

during a period of 3 to 4 days. Each branch was enclosed with a fine-mesh cloth sleeve cage which served as an<br />

Mattson, W.J., NiemelS., R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

86


enclosure %r 20 post-diapausing second instar larvae. Larvae were placed in the cages when lO0°C-days were attained, <strong>or</strong><br />

approximately 2 weeks pri<strong>or</strong> to opening of vegetative buds,<br />

Sample branches were cut when budw<strong>or</strong>ms reached pupal stage. Pupae were weighed, moths were melted, the<br />

number and the weight of eggs laid were rec<strong>or</strong>ded every day during the oviposition period. The progeny were placed at 2°C<br />

f<strong>or</strong> a 29 week long diapause period, and the survival of the progeny was rec<strong>or</strong>ded at the end of diapause. On each branch, the<br />

weight of current-year foliage produced by the tree, and the weight of current-year and 1-year old foliage removed by the<br />

larvae were estimated.<br />

Data were analyzed using the Statistical Analysis System (Sas Institute 1988). N<strong>or</strong>mality and variance homogeneity<br />

tests were perf<strong>or</strong>med bef<strong>or</strong>e data were subjected to a nested-fact<strong>or</strong>ial, multivariate analysis of variance with individual trees<br />

nested within flowering class and crown sections crossed with flowering class. Percent survival was analyzed using Chisquare<br />

analysis.<br />

Lab<strong>or</strong>at<strong>or</strong>y Rearing Experiment<br />

The impact of spruce budw<strong>or</strong>m feeding behavi<strong>or</strong> on the effects of flowering on the insect growth and development<br />

was determined using lab<strong>or</strong>at<strong>or</strong>y rearing experiments with fresh food from the field. Also, <strong>this</strong> approach allowed us to<br />

separate the food effect from the microhabitat effect provided by the staminate flower clusters.<br />

A total of eight feeding scenarios were simulated under lab<strong>or</strong>at<strong>or</strong>y conditions (T = 12°C, RH = 65%, 18L:6D) using<br />

the methodology of Bauce et al. (1994). These conditions c<strong>or</strong>responded to the average conditions prevailing at the experimental<br />

field site during the period of spruce budw<strong>or</strong>m feeding activity. The eight feeding scenarios were (1) larvae fed on<br />

staminate flowers during 4 days, and current-year foliage from the midcrown of flowering trees thereafter, (2) larvae mined<br />

1-year old needles until vegetative budbreak, and fed on current-year foliage from the lower crown section of flowering trees<br />

thereafter, (3) larvae fed on staminate flowers during 4 days, current-year foliage from the midcrown of flowering trees until<br />

they reached their larval sixth instar, and 1-year old foliage from the midcrown of flowering trees thereafter, (4) larvae fed on<br />

staminate flowers during 4 days, current-year foliage from the midcrown of flowering trees until they reached their larval<br />

sixth instar, and current-year foliage from the lower section of flowering trees thereafter, (5) larvae mined 1-year old needles<br />

until vegetative budbreak, and fed on current-year foliage from the midcrown of nonflowering trees thereafter, (6) larvae<br />

mined l-year old needles until vegetative budbreak and fed on current-year foliage from the lower crown section of nonflowering<br />

trees thereafter, (7) larvae mined l-year old needles until vegetative budbreak, fed on current-year foliage from the<br />

midcrown of nonflowering trees until they reached their larval sixth instar, and 1-year old foliage from the midcrown of<br />

nonflowering trees thereafter, and (8) larvae mined 1-year-old needles until vegetative budbreak, fed on current-year foliage<br />

from the mid-crown of nonflowering trees until they reached their larval sixth instar, and current-year foliage from the lower<br />

crown section of nonflowering trees thereafter. The scenarios 3, 4, 7, and 8 simulated a lack of current-year foliage when<br />

larvae reach sixth instar. This phenomenon usually occurs when larval density is high. Scenarios 3 and 7 simulated a<br />

backfeeding (feeding on 1-year old needles) while scenarios 4 and 8 simulated dispersal of the population from the midcrown<br />

to the lower crown section. Backfeeding and dispersal to the lower crown section were observed by Blais (1952) at high<br />

budw<strong>or</strong>m population density.<br />

Two trees served as food source f<strong>or</strong> two groups of 25 individually reared female larvae per feeding scenario. The<br />

food was replaced in the rearing containers at 2-d intervals and budw<strong>or</strong>m growth and development were rec<strong>or</strong>ded f<strong>or</strong> each<br />

instar separately. Insect development was monit<strong>or</strong>ed 24 hours a day every 6 hours and every 15 min during the moult. One<br />

hour after molting from one instar to the next, bef<strong>or</strong>e feeding was resumed, each larvae was weighted along with he newly<br />

molted larval skin and then transferred to a new twig. Pupae were weighed 8 hours after pupation. N<strong>or</strong>mality and variance<br />

homogeneity tests were perf<strong>or</strong>med bef<strong>or</strong>e data were subjected to a two-stages nested multivariate analysis of variance with<br />

individual trees nested within feeding scenarios.<br />

87


RESULTS<br />

Field Rearing Experiment<br />

The production of staminate flox_ers significantly affected spruce budw<strong>or</strong>m larval biology (9: Wilk's lambda F[2,171<br />

= 18.3, p = 0.0(_1, if: Wilk's lambda F[2,171 = 13. l, p = 0.0004) (Fig. 1, 2). Larval biolog 5 also significantly xaried<br />

acc<strong>or</strong>ding to crown sections (9: Wilk's lambda F[2,17] = 458.6, p = 0.0001 ; Of: Wilk's lambda F:[2,171= 331.8, p = 0.0001).<br />

However, the interaction between flo_vering class and crown section was significant (9: Wilk's lambda F[2,17] = 329.3, p =<br />

0.0(X)I o": Wilk's lambda F[2,17] = 549.9, p = 0.0001). This significant interaction results from the fact that floxvering<br />

affected budw<strong>or</strong>m biology only in the midcrown of flowering trees where staminate flowers were available f<strong>or</strong> consumption<br />

(Figs. 1, 2).<br />

F F NF NF<br />

o_ a a<br />

E loo ab<br />

,,,,,,.-1<br />

..c:<br />

.__ 80 a<br />

.c<br />

0"J<br />

a<br />

-<br />

)<br />

7<br />

a<br />

o<br />

.._<br />

Ck<br />

60<br />

-'i<br />

[1.. 40<br />

2o<br />

o ,,,,, ,i =,.,.<br />

Figure 1.--Pupal weight of male and female spruce budw<strong>or</strong>m reared in the midcrown and the lower sections (*) of flowering<br />

(F) and nonflowering (NF) balsam fir trees. Staminate flowers were present only in the midcrown of flowering trees<br />

(n = 10 trees). F<strong>or</strong> a given sex, means followed by the same letter are not significantly different (p < 0.05).<br />

Both male and female budw<strong>or</strong>ms that had access to staminate flowers in the midcrown of flowering trees exhibited 5<br />

d sh<strong>or</strong>ter development time (9: F[1,18] = 32.7, p = 0.0001; _: F[1,18] = 27. l, p = 0.0001) and 9_'_(9) to 15% (o_) reduced<br />

pupal weight (9: F[1,18] = 12.3, p = 0.0026; if: F[1,1811= 4.9, p = 0.0397) compared to those located where staminate<br />

flowers were not available f<strong>or</strong> consumption. However, no significant impact of flowering on budw<strong>or</strong>m survival was detected<br />

(Chi-square = 75, df = 69, p = 0.28, n = 40) (Fig. 3).<br />

Insects reared on flowering branches had lower fecundity (F[ 1,18] = 10.8, p = 0.0041) than those reared on nonflowering<br />

branches (Fig. 4), but they produced individual eggs of similar weight (F[1,8] = 1.8, p = 0.21) (Fig. 5). At the end of<br />

the 29 week long diapause period, no significant difference was detected in terms of survival of the progeny maintained at<br />

2°C during their diapausing stage (Chi-square = 132, df = 129, p = 0.41, n = 46) (Fig. 6).<br />

Flowering branches produced significantly less current-year R_liagethan nonflowering branches (F[1,181 = 4.4, p =<br />

0.04) (Fig. 7). Although the insects reared on flowering branches removed approximately the same amount of current-year<br />

foliage as those reared on nonflowering branches (F[1,1811= 1.95, p = 0.18), they back-fed on 1-year old foliage, while the<br />

others did not (F[l, 18] = 23, p = 0.0(Y01) (Fig. 7). This indicates a possible reduction in the nutritive quality of the currentfoliage<br />

produced by flowering branches.<br />

88


2,500<br />

F<br />

000<br />

F NF NF<br />

b _7 a _;_ a _7 a<br />

E<br />

EL<br />

i •<br />

__0<br />

¢_1,000<br />

i 1 ; ; i<br />

• !<br />

o i i ,<br />

| ,<br />

I ,<br />

Figure 2.--Development time from post diapausing second instar to moth emergence of male and female spruce budw<strong>or</strong>m<br />

reared in the midcrown and the lower sections (*) of flowering (F) and nonflowering (NF) balsam fir trees. Stami-<br />

nate flowers were present only in the midcrown of flowering trees (n = 10 trees). F<strong>or</strong> a given sex, means followed by<br />

the same letter are not significantly different (p < 0.05).<br />

6o<br />

a<br />

F F NF NF<br />

t_<br />

_o _ _ _<br />

20 :<br />

lO [ EE<br />

o.....,....... _/_ ................... _, .....Z_ , I , b,-<br />

Figure 3.mPercent survival of spruce budw<strong>or</strong>m reared in the midcrown and the lower sections (*) of flowering (F) and<br />

nonflowering (NF) balsam fir trees. Staminate flowers were present only in the midcrown of flowering trees (n = 10<br />

trees). Means followed by the same letter are not significantly different (p < 0.05).<br />

' i<br />

89


50+ + +<br />

"o a<br />

-_ 7 a<br />

° a _7<br />

°_ __<br />

o b . i ;<br />

1 . , i<br />

: :" ! [<br />

c I I<br />

50 i : I<br />

o , I<br />

o , / , ,t , J/ , .Z_<br />

Figure 4.mFecundity of spruce budw<strong>or</strong>m reared in the midcrown and ihe lower sections (*) of flowering (F) and nonflowering<br />

(NF) balsam fir trees. Staminate flowers were present onlyin the midcrown of flowering trees (n = 10 trees).<br />

Means followed by the same letter are not significantlydifferent (p < 0.05).<br />

E<br />

0.35<br />

F F NF NF<br />

A & & &<br />

U<br />

v a<br />

03 a a ./'---"-2<br />

0"} 0.25 L I<br />

_o_ / /<br />

._><br />

"O t'-<br />

.m<br />

N--.<br />

o o.15 i<br />

..c:: !<br />

.m03<br />

E<br />

(D E<br />

o.1 iI ¢) t<br />

0.05 f F<br />

J<br />

o , / , Z , _ , /<br />

Figure5.--Mean weightof individual eggslaid by sprucebudw<strong>or</strong>mrearedin themidcrownandthelowersections(*) of<br />

flowering (F) and nonflowering (NF) balsam fir trees. Staminateflowers werepresent only in the midcrown of<br />

flowering trees (n = 10 trees). Means followed by the same letter are not significantly different (p < 0.05).<br />

90<br />

i


8O<br />

D<br />

09 i<br />

4,'.)<br />

F AS4L<br />

a<br />

Z___7<br />

2o !<br />

i<br />

,<br />

.i J J<br />

Figure 6.--Percent survival at the end of the diapause period of the progeny of spruce budw<strong>or</strong>m reared in the midcrown and<br />

the lower sections (*) of flowering (F) and nonflowering (NF) balsam fir trees. Staminate flowers were present only<br />

in the midcrown of flowering trees (n = I0 trees). Means followed by the same letter are not significantly different (p<br />

< 0.05).<br />

20<br />

.+.,a<br />

x2 a<br />

-cJ<br />

o?<br />

.N<br />

-6 b<br />

LL a<br />

5 a<br />

a<br />

I<br />

0 ! ! ! .... - ......<br />

D defoliation<br />

D foliage production<br />

backfeeding<br />

Figure 7._Weight of current-year foliage produced by flowering (F) and nonflowering (NF) balsam fir trees at two crown<br />

levels (*), and weight of current-year and l-year old (backfeeding) foliage consumed by spruce budw<strong>or</strong>m reared on<br />

flowering and nonflowering balsam fir trees at two crown levels (*). Staminate tlowers were present only in the<br />

midcrown of flowering trees (n = 10 trees). F<strong>or</strong> a given parameter, means followed by the same letter are not<br />

significantly different (p < 0.05).<br />

I<br />

91


Lab<strong>or</strong>at<strong>or</strong>y Rearing Experiment<br />

P,esults from the lab<strong>or</strong>at<strong>or</strong>y rearing experiment indicated that the various feeding scenarios tested in <strong>this</strong> study<br />

significantly affected female spruce budwonn growth (Fig. 8) and development (Fig. 9) (Wilk's lambda F[49,15] = 4.7, p =<br />

().(}012).<br />

2OO<br />

st'5 [_ st'6 [_ pupa<br />

''O ti g<br />

D3<br />

E<br />

_ loo diE]7 t _ ;_<br />

[i ........ _ j I !<br />

5O<br />

_[<br />

,<br />

,r<br />

i !<br />

[<br />

i<br />

'<br />

t<br />

; I<br />

[<br />

i<br />

',<br />

] I<br />

l<br />

!<br />

a<br />

!<br />

a , ] a :/ a /_ I a / a /<br />

b _ _ b / b / t i ,/ / --"-<br />

' " I' // /"<br />

..__i/ i.J=. ,, / ,,a/,, IJa _a...g..,Z_i /<br />

Figure 8.---Female spruce budw<strong>or</strong>m weight per instar acc<strong>or</strong>ding to various feeding scenarios using flesh food from flowering<br />

(F) and from nonflowering (NF) balsam fir trees. Two trees served as food sources f<strong>or</strong> two groups of 25 individually<br />

reared larvae per feeding scenario (T = 12°C, RH = 65%, 18L:6D, + larvae fed on current-year foliage from the given<br />

crown section, - larvae fed on current-year foliage pri<strong>or</strong> to reaching sixth instar and l-year old foliage thereafter, *<br />

larvae fed on current-year foliage from the midcrown pri<strong>or</strong> to reaching sixth instar and current-year foliage from the<br />

lower"crown section thereafter). F<strong>or</strong> a given instar, means followed by the same letter are not significantly different<br />

(p < 0.05).<br />

Insects fed on pollen during their early stages of development were smaller when they reached their sixth instar than<br />

those that did not have access to staminate flowers (Fig. 8). However, they had similar pupal weight as those that did not eat<br />

pollen but had access to current-year foliage during their whole development. The presence of pollen in the insect diet<br />

caused a significant 5 d reduction in development time pri<strong>or</strong> to reaching sixth instar, and a significant 3 d reduction during<br />

pupal stage (Fig. 9). Although backfeeding during the insect sixth instar did not significantly affect development time, it<br />

reduced the pupal weight of larvae fed on food from flowering and nonflowering trees by 32% and 28% respectively (Fig. 8).<br />

On the other hand, dispersal from the midcrown to the lower crown section, when larvae reached sixth instar, did not significandy<br />

affect pupal weight. However, <strong>this</strong> feeding scenario resulted in a 6 d increase in the development time from sixth<br />

instar to moth emergence (Fig. 9).<br />

92<br />

I


2,500 _ st2-st4 El st5 0 st6 [_ pupa<br />

,ooo<br />

"-" b b b bc<br />

0.> --m 7 bc _ "I"--7 .----.<br />

:_ 1,5oo d d _ _..a..4<br />

@ 7 ...... -8,,'q /-'-7 _ .....-," ...... l".b',.,_ ...... a.," an',._<br />

E ....._'1 ' "t '\,a\. : x ._<br />

_9_O >: .,. _t ._ - - i ._-.I "


development time that they obtained from feeding pollen during their early stages of developrnent. On the other<br />

tland, when old larvae dispersed to the lower crown section, they avoided the negative effects of backfeeding, but they lost<br />

the advantage in development time that they obtained from feeding on pollen. Results from <strong>this</strong> study indicated that the<br />

production of staminate flowers by balsam fir trees could have opposite effects on spruce budw<strong>or</strong>m population dynamics<br />

depending upon the insect population density when flowering occurs.<br />

ACKNOWLEDGMENTS<br />

We wish to thank M. Charest, M. Crdpin, D. Verge, G. M<strong>or</strong>eau, J.F. Demers, S. Perreault, C. Paquet, I. C<strong>or</strong>d, J.<br />

F<strong>or</strong>est, L. Roy, and M. Marin f<strong>or</strong> technical assistance during insect rearings. Insects were supplied by the Insect Production<br />

Unit of the F<strong>or</strong>est Pest Management Institute, Sault Ste. Marie, Ontario, Canada. Financial supp<strong>or</strong>t came from the Natural<br />

Sciences and Engineering <strong>Research</strong> Council of Canada.<br />

LITERATURE CITED<br />

BATZER, H.O. and JENNtNG, DT. 1980. Numerical analysis of a jack pine budw<strong>or</strong>m outbreak .in various densities of jack<br />

pine. Environ. Entomol. 9: 514-524.<br />

BAUCE, E. 1986. I_tude des variations de teneur en fibre brute du feuillage de sapin baumier, Abies bals'ames (L.) Mill.,<br />

induites par diffErents facteurs de stress et leurs implications sur la t<strong>or</strong>deuse des bourgeons de l'dpinette,<br />

Ch<strong>or</strong>istoHeurafumiferana (Clem.). M.Sc, thesis, Universit6 Laval, Ste-Foy (Quebec), Canada, 54 p.<br />

BAUCE, E., CREPIN, M., and CARISEY, N. 1994. Spruce budw<strong>or</strong>m growth, development and food utilization on young<br />

and old balsam fir trees. Oecologia (in press).<br />

BLAIS, JR. 1952. The relationship of the spruce budw<strong>or</strong>m to the flowering condition of balsam fir. Can. J. Zoot. 30:1-29.<br />

BELANGER, L., DUCRUC, J.R, and PINEAU, M 1983. Analyses et commentaires. Proposition d'uhne mdthodologie<br />

d'inventaire dcologique adaptde au territoire f<strong>or</strong>estier pdriurbain. Nat. Can. 110: 459-476.<br />

GRAHAM, S.A. 1935. The spruce budw<strong>or</strong>m on Michigan pine. University of Michigan, School of F<strong>or</strong>est Conservation,<br />

Bull No 6.<br />

GRANDTNER, MM. 1966. La vdgdtation fbrestiare du Qudbec mdridional. Universitd Laval, Ste-Foy (Qudbec), Canada.<br />

GREENBANK, D.O. 1963. S taminate flowers and the spruce budw<strong>or</strong>m. Mem. Entomol. Soc. Can. 31: 202-218.<br />

JAYNES, H.A. and SPEERS, C.F. 1949. Biological and ecological studies of spruce budw<strong>or</strong>m. J. Econ. Entomol. 42: 221-<br />

225.<br />

LEJEUNE, R.R. 1950. The effect of jack pine staminate flowers on the size of larvae of the jack pine budw<strong>or</strong>m. Can. Ent.<br />

82: 34-43.<br />

LEJEUNE, R.R. and BLACK, W.E 1950. Populations of larvae of the jack pine budw<strong>or</strong>m. F<strong>or</strong>. Chron. 26: 152-156.<br />

MATTSON, W.J., HAACK, R.A., LAWRENCE, R.K., and SLOCUM, S.S. 1991. Considering the nutritional ecology of<br />

spruce budw<strong>or</strong>m in its management. F<strong>or</strong>. Ecol. Manage. 39: 183-2110.<br />

MILLER, C.A. 1957. A technique f<strong>or</strong> estimating the fecundity of natural populations of spruce budw<strong>or</strong>m. Can. J. Zool. 35:<br />

1-13.<br />

SAS INSTITUTE. 1988. SAS user's guide; statistics. 1988. SAS Institute Inc., Cary, NC.<br />

95


96<br />

SLANSKY, E, Jr. 1990. Insect nutritional ecology as a basis f<strong>or</strong> studying host plant resistance. Fl<strong>or</strong>ida Entomol. 73" 360-<br />

378.<br />

VOLNEY, W.J.A. 1988. Analysis of hist<strong>or</strong>ic jack pine budw<strong>or</strong>m in the Prairie province of Canada. Can. J. F<strong>or</strong>. Res. 18:<br />

1152-1158.<br />

WELLINGTON, W.G. 1950. Effects of radiation on the temperatures of insectan habitats. Sci. Agr. 30: 209-234.


STAMINATE FLOWERING AND TREE PHENOLOGY AFFECT THE<br />

PERFORMANCE OF THE SPRUCE BUDWORM<br />

WILLIAM J. MATTSON, BRUCE A. BIRR, and ROBERT K. LAWRENCE<br />

<strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong><br />

Pesticide <strong>Research</strong> Center, Michigan State University, E. Lansing, M148824, USA<br />

INTRODUCTION<br />

The imp<strong>or</strong>tant role of staminate flowers in the ecology and population dynamics of spruce budw<strong>or</strong>m has been the<br />

subject of speculation f<strong>or</strong> m<strong>or</strong>e than 50 years (Blais 1952, Greenbank 1963). This suspicion has been fueled by the fact that<br />

high populations of budw<strong>or</strong>m have hist<strong>or</strong>ically been associated with mature, flower-bearing trees (Mott 1963). M<strong>or</strong>eover,<br />

overwintering budw<strong>or</strong>m larvae typically spin their silk shelter in the old staminate flower bracts, and spring emerging 2nd<br />

instars prefer to mine and feed in staminate flower buds rather than old needles. Budw<strong>or</strong>ms emerge 1 to 4 weeks bef<strong>or</strong>e<br />

vegetative buds have begun to expand and are suitable f<strong>or</strong> attack, whereas flower buds, when present, are always m<strong>or</strong>e<br />

advanced and thus better synchronized with the early, vernal budw<strong>or</strong>ms. Hence, second-stage larvae are invariably concentrated<br />

in the host's expanding flower buds. On the other hand, whenever budw<strong>or</strong>m emergence is retarded relative to tree<br />

phenology, then larvae tend to go directly to expanding foliage rather than the staminate flower buds (Greenbank 1963).<br />

Obviously, then, the imp<strong>or</strong>tance of flower buds to budw<strong>or</strong>ms may be linked to the phenology of the hosts' vegetative buds<br />

relative to the timing of spring emergence of the budw<strong>or</strong>ms.<br />

Larvae that feed within staminate flower clusters may actually grow faster than foliage-feeding larvae because such<br />

clusters are miniature greenhouses, trapping m<strong>or</strong>e radiant energy than comparable pure foliage environments. However,<br />

there is little other experimental evidence to supp<strong>or</strong>t the hypothesis that staminate flowers can enhance budw<strong>or</strong>m perf<strong>or</strong>mance<br />

and thus contribute to outbreak development. F<strong>or</strong> example, Jaynes and Speers (1949) and Blais (1952) rep<strong>or</strong>ted that<br />

there is no difference in fecundity of budw<strong>or</strong>ms having fed on pollen <strong>or</strong> new foliage.<br />

Herms and Mattson (1992) hypothesized that plant reproduction can also indirectly affect herbiv<strong>or</strong>e perf<strong>or</strong>mance<br />

through its strong impact on the allocation of plant nutrients and photosynthates. When plants produce large crops of pollen<br />

and/<strong>or</strong> seeds, the reproductive <strong>or</strong>gans may preferentially receive scarce nutrients and energy that might otherwise have gone<br />

to plant defenses <strong>or</strong> even new growth on which phytophages depend. Hence, staminate flowering could indirectly impact<br />

budw<strong>or</strong>ms by reducing the concentrations of nutrients and perhaps even altering the levels of secondary metabolites in plant<br />

foliage.<br />

This study was undertaken to simulate and expl<strong>or</strong>e the consequences of abundant staminate flower production on the<br />

perf<strong>or</strong>mance of pre-outbreak spruce budw<strong>or</strong>m populations. We hypothesized that budw<strong>or</strong>m larval perf<strong>or</strong>mance would<br />

depend not just on the availability of staminate flower but also on the phenology of a tree's vegetative buds. We predicted<br />

that the presence of staminate flowers would be most imp<strong>or</strong>tant on very late flushing trees, and least imp<strong>or</strong>tant on very early<br />

flushing trees.<br />

METHODS<br />

In 1989 and 1991 we selected 20 half-sibling balsam fir, Abies balsamea, that were flowering at the Kellogg Experi-<br />

mental F<strong>or</strong>est (Michigan State University) near Augusta, Michigan. Trees were planted in 1970 and about 6-8 m tall at the<br />

time of the experiments. Bef<strong>or</strong>e budbreak (April 21, 1989, and April 11, 1991), we selected per tree two flowering and two<br />

nonflowering branches near midcrown, which were then enclosed with a fine-mesh, cloth sleeve cage wherein we placed<br />

about 20 ready to emerge second instar budw<strong>or</strong>m larvae still in their silken hibernacula on a gauze patch. Exactly 1 week<br />

Mattson, W.J., Niemel_i, P.,and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

97


later, we selected one m<strong>or</strong>e flowering and nonflowering branch on each tree which we once again enclosed with sleeve cages<br />

arid ir_;e_ted about 20 second-stage budw<strong>or</strong>ms. These two batches came from the same population, and they were randomly<br />

allocated to treatments. Hereafter, we respectively refer to them as coh<strong>or</strong>t 1 and coh<strong>or</strong>t 2. Approximately 2 weeks later we<br />

removed the gauze patches and counted the number of larvae which failed to emerge from their overwintering hibernacula.<br />

This gave us the initial number of insects in each sleeve cage. When most insects had reached the pupal stage, cages were<br />

removed by cutting each branch at its base. We then searched each bag f<strong>or</strong> budw<strong>or</strong>m larvae and pupae. Pupae were st<strong>or</strong>ed<br />

immediate[y in plastic vials and checked daily until all adults emerged. Moths were first frozen, then oven dried to constant<br />

weight and weighed and sexed. Because male and females responded in the same way to treatments, we pooled sexes by first<br />

converting males to female equivalents. Each male observation was multiplied by a constant (k) that was derived from the<br />

grand meaas (gdm) of the particular variable in question (e.g., weight, developmental time, developmental rate, etc): k_= gdm<br />

fcmate/gdm male, f<strong>or</strong> each coh<strong>or</strong>t and year. We calculated survival per sleeve cage by summing all larvae that achieved the<br />

final 6th larval stage (even though some were parasitized <strong>or</strong> preyed upon by stink bugs), pupae, and adults, and divided <strong>this</strong><br />

grand total by the mm_berof larvae that actually emerged from their hibernacula.<br />

In 1991, we selected 32 (16 flowering, 16 nonflowering) white spruce trees, Picea glauca, that were growing at the<br />

edge of a provenance plantation (<strong>USDA</strong> N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong>) near Wellston, Michigan. Trees were<br />

planted in 1963, and were approximately 8-10 m tall at the time of the experiment. Just as above we selected (on May 3) two<br />

branci_es at midcrown level from each the flowering and the nonflowering trees f<strong>or</strong> enclosure with sleeve cages and budw<strong>or</strong>ms.<br />

We placed only one coh<strong>or</strong>t of budw<strong>or</strong>m larvae on these trees, however. We collected and counted surviving budw<strong>or</strong>ms<br />

when most had reached the pupal stage.<br />

Aft trees were sc<strong>or</strong>ed f<strong>or</strong> their phenological development acc<strong>or</strong>ding to the protocol of Nienstaedt and King (l 970) at<br />

the time of placing insects on the trees.<br />

Difference between coh<strong>or</strong>ts 1 and 2 in the balsam fir study were first tested via a 2 x 2 fact<strong>or</strong>ial, randomized block<br />

design where trees were treated as blocks and flowers and coh<strong>or</strong>t timing were treated as treatments. Next each coh<strong>or</strong>t was<br />

analyzed separately to test specifically f<strong>or</strong> flower by tree phenology effects on budw<strong>or</strong>m perf<strong>or</strong>mance (weight gain, growth<br />

rate/day, days to complete development, and survival to the pupal stage). We subjected the data to univariate anova after<br />

appropriate transflmnations of the data (i.e., arcsin (%survival)). The balsam fir study was analyzed as a completely randomized,<br />

spiit,..piot design with phenology (3 classes) treated as the main plot and flowering (2 levels) as the subplot. The white<br />

spruce study was analyzed as a completely randomized 3 x 2 (phenology x flowering) fact<strong>or</strong>ial.<br />

Larval Survival<br />

RESULTS<br />

Balsam Fir: Flowering and Phenology Effects on Budw<strong>or</strong>m<br />

First coh<strong>or</strong>t survival was clearly affected by both flowering (F) and flowering x tree phenology (P) classes (Table 1).<br />

[n 1989, the significant F x P interaction was due largely to the enhancing effect of flowering, as expected, on the Sateflushing<br />

trees, where survival averaged 65.3% with, and 43.2% without flowers (Fig. 1). The smallest flowering effect<br />

occurred on early flushing tree (57.0% vs 53.6%, with and without flowers, respectively). In 1991, even though there were<br />

no significant main <strong>or</strong> interactio_ effects (Table 2),the largest positive impact of flowering was evident on the late flushing<br />

trees (as in 1989) where survival averaged 48.2% with, and 40.6% without flowers (Fig. 1).<br />

Second coh<strong>or</strong>t survival in 1989 was significantly (p < 0.02) enhanced by flowering but not affected by phenology, n<strong>or</strong><br />

by the F x P interaction (Table 1). In 1991, there were again significant flower, and F x Perfects (Table 2). This was due<br />

primarily to the huge difference in survival rates between flowering (54.1%) and nonflowering (32.0%) branches on late<br />

flushing trees (Fig. I).<br />

Testing the grand means f<strong>or</strong> coh<strong>or</strong>ts I and 2 in 11989and 1991 revealed in both cases that survival was higher f<strong>or</strong> the<br />

first coh<strong>or</strong>ts, but only statistically significant in 1989 (55% vs 43%).<br />

98


Table 1.--Mean spruce budw<strong>or</strong>m perf<strong>or</strong>mance (female equivalents) on balsam fir in 1989 on early, middle and late<br />

phenology trees (main plots) and flowering and nonflowering branches (subplots). Means varying significantly<br />

(p < 0.05) among main effects have different letters. F<strong>or</strong> coh<strong>or</strong>t one, mean survival was effected by flowering<br />

and by,the flowering x phenology interaction. F<strong>or</strong> coh<strong>or</strong>t two, mean survival was affected by flowering.<br />

Budw<strong>or</strong>m perf<strong>or</strong>mance Trees (main Elots)--Phenolo_ class Subplot--Flowerin i class<br />

variable Early Middle Late Yes No<br />

Coh<strong>or</strong>t one<br />

Survival (larvae) 0.55 0.57 0.54 0.62a 0.49b<br />

Weight dwt (rag) 23.93 22.75 22.44 23.18 22.72<br />

Dev. time (days) 52.94 53.37 52.93 52.91 53.34<br />

Growth rate (mg/da) 0.45 0.43 0.43 0.44 0.43<br />

Coh<strong>or</strong>t two<br />

Survival (larvae) 0.43 0.42 0.45 0.47a 0.39b<br />

Weight dwt (rag) 26.11 23.58 24.23 24.53 24.1<br />

Dev. time (days) 50.45 51.27 50.88 51. | 1 50.84<br />

Growth rate (mg/da) 0.55 0.46 0.48 0.48 0.48<br />

0.70 _- 1989<br />

0.65 _ _,ow_.,[] N_,,<br />

0.60<br />

t Coh<strong>or</strong>t 1<br />

Figure 1.---Survival rates of two "_ ,//_<br />

coh<strong>or</strong>ts of spruce budw<strong>or</strong>m _- #/'A<br />

larvae in 1989 (upper panel) ._ 0.50 /)///_<br />

and 1991 (lower panel) on ?' :.:;,:.:,:'.:il<br />

flowering and nonflowering _ 0.4.5 ',:,::'.,.::,:.:_<br />

branches on 20 balsam fir trees ]<br />

classes: early, mid and late 0.35<br />

flushing at the Kellogg<br />

divided into three phenology<br />

Experimental F<strong>or</strong>est of<br />

o.ao f<br />

0.30 --<br />

_i1<br />

Michigan State University in<br />

Augusta, Michigan. Flowering 0.55 l<br />

0.55f i] .. ,, oo o,,<br />

effects were statistically E2:I_,o_,,,,,_ s_,_ coh<strong>or</strong>ta<br />

significant (p < 0.05) in one <strong>or</strong> 0.50<br />

both coh<strong>or</strong>ts of both years<br />

(Tables 1 and 2).<br />

:_',_:!:::<br />

¢ Coh<strong>or</strong>t<br />

and flowering x phenology t| 1981 .:,,:.:,',.':.':,_,_i!<br />

o.4o<br />

0.35<br />

0.30 __ __dIZ<br />

early- I mid- 1 late- 1 early-2 mid-2 late-2<br />

Tree Dhenolo_ty & budw<strong>or</strong>m coh<strong>or</strong>t classes<br />

L///<br />

V//<br />

V'//<br />

99


Table 2.---Mean spruce budw<strong>or</strong>m perf<strong>or</strong>mance (female equivalents) on balsam fir in 1991 on early, middle and late phenology<br />

trees (main plots) and flowering and nonflowering branches (subplots). Means varying significantly (p < 0.05)<br />

among main effects have different letters. F<strong>or</strong> coh<strong>or</strong>t one, mean weight and growth rate were significantly affected<br />

by phenology, and mean development time by the f x p interaction. F<strong>or</strong> coh<strong>or</strong>t 2, only mean survival was significantly<br />

affected by flowering, and by the flowering x phenology interaction.<br />

Budw<strong>or</strong>m perf<strong>or</strong>mance ..... Trees, (main plots)--Phenology class SubplotmFlowerin_g "class<br />

variable Early Middle Late Yes No<br />

Coh<strong>or</strong>t one<br />

Survival (larvae) 0.41 0.41 0.44 0.42 0.42<br />

Weight dwt (rag) 23.41b 26.47a 25.72a 24.79 25.00<br />

Dev. time (days) 51.29 50.97 51.05 51.05 51.23<br />

Growth rate (mg/da) 0.46b 0.52a 0.50a 0.49 0.49<br />

Coh<strong>or</strong>t two<br />

Survival (larvae) 0.45 0.47 0.43 0.47a 0.4 l b<br />

Weight dwt (rag) 24.37 25.27 25.53 24.51 25.44<br />

Dev. time (days) 46.02 46.16 45.65 45.92 45.95<br />

Growth rate (mg/da) 0.51 0.55 0.56 0.53 0.55<br />

Larval Growth<br />

There were no significant flower <strong>or</strong> F x P effects on weight gain (mg dwt) by the first <strong>or</strong> second budw<strong>or</strong>m coh<strong>or</strong>t in<br />

either 1989 <strong>or</strong> 1991 (Tables 1,2). However, there was a significant phenology effect on coh<strong>or</strong>t 1 in 1991, where middle and<br />

late flushing trees yielded bigger adults than early flushers (26.5 vs 23.4 mg dwt).<br />

Likewise, there were no significant flowering, <strong>or</strong> F x P effects on budw<strong>or</strong>m growth rates (mg/da) f<strong>or</strong> either coh<strong>or</strong>ts in<br />

1989, and 1991 (Table 1, 2). However, there was a significant phenology effect on the growth rate of the first 1991 coh<strong>or</strong>t.<br />

Budw<strong>or</strong>ms on middle and late trees grew faster than those on early flushing trees (0.52 mg/da vs 0.46 mg/da). There was<br />

also a significant F x P effect on development time f<strong>or</strong> coh<strong>or</strong>t one (Fig. 2), but <strong>this</strong> may have been spurious.<br />

Comparing the grand means f<strong>or</strong> coh<strong>or</strong>ts 1 and 2 in 1989revealed that growth (22.95 vs 24.31 mg) was significantly<br />

higher and development sh<strong>or</strong>ter (53.13 vs 50.97 da) f<strong>or</strong> the second than first coh<strong>or</strong>t. By contrast, in 1991, there were no<br />

significant differences in overall growth between the two coh<strong>or</strong>ts, but development time was, as bef<strong>or</strong>e, much sh<strong>or</strong>ter f<strong>or</strong> the<br />

second (51.14 vs 45.93 da), causing its growth rates to be significantly higher (0.49 vs 0.54 mg/da).<br />

Survival<br />

White Spruce: Flowering and Phenology Effects on Budw<strong>or</strong>m<br />

Flowering and phenology had no significant effects on survival of larvae (Table 3), but the F x P interaction was<br />

nearly, significant (p < 0.10). There was an apparent trend f<strong>or</strong> survival to increase with increasingly retarded phenology on<br />

nonflowering trees, and the opposite trend on flowering trees (Fig. 2). The fact that the two trends run counter to each<br />

explains why the main effects of flowering and phenology were insignificant. The grand means from flowering and nonflowering<br />

trees in each phenology class canceled one another out.<br />

Larval Growth<br />

There were strong, consistent phenology effects on both total weight gain and growth rates, the pattern being identical<br />

on both flowering and nonflowering trees (Table 3). Budw<strong>or</strong>m mass and average growth per day increased with increasingly<br />

retarded tree phenology, Insects on the late flushing trees averaged 13.5% larger than those from the early flushing trees<br />

(32.5 mg vs 28.6 rag). Likewise their daily growth rates were 14.5% higher (80 mg/da vs 70 mg/da). Neither flowering, n<strong>or</strong><br />

t00


52,00<br />

51,75 -<br />

51,50 -<br />

51.25<br />

50.75 _///<br />

50.00<br />

0,70<br />

0.65<br />

0.60<br />

0.55<br />

OA5<br />

0.50I<br />

0.40 _-<br />

0.35<br />

38<br />

_!ow_rs _ None<br />

36 - _Z] _=lower__ No


Table 3..--Mean spruce budw<strong>or</strong>m perf<strong>or</strong>mance (female equivalents) on white spruce in 1991 on early, middle, and late<br />

phenology trees and flowering and nonflowering trees. Means varying significantly (p < 0.05) due to main effects<br />

have different letters. Only mean weight and development time were affected by phenology. There were no flowering<br />

effects.<br />

Budw<strong>or</strong>m perf<strong>or</strong>mance Tree phenol0_y class Tree flowering, class<br />

variable Early Middle Late Yes No<br />

Coh<strong>or</strong>t one<br />

Survival (larvae) 0.57 0.57 0.55 0.57 0.57<br />

Weight dwt (rag) 28.61b 30.96a 32.48a 30.13 311.33<br />

Dev. time (days) 41.11 40.93 40.85 40.96 40.95<br />

Growth rate (mg/da) 0.70b 0.76a 0.80a 0.74 0.77<br />

F x P had any significant (p < 0.05) effects on growth, development time, <strong>or</strong> growth rates. Although, there was a hint (p <<br />

0.08) of a tendency f<strong>or</strong> nonflowering trees to be superi<strong>or</strong> f<strong>or</strong> larval perf<strong>or</strong>mance than flowering trees (Fig. 2).<br />

DISCUSSION and CONCLUSIONS<br />

Yearly Differences in Protocol and Weather Weaken the Tests<br />

The differences in the effects of flowering and phenology on spruce budw<strong>or</strong>m perf<strong>or</strong>mance between years is ahnost<br />

certainly due to the differences in host tree phenology when the studies were initiated. In 1989, we placed our first coh<strong>or</strong>t of<br />

budw<strong>or</strong>ms on the trees at 133 degree days (dd), very close to the time native budw<strong>or</strong>m populations would have been emerging<br />

(est. at 100 dd). The second coh<strong>or</strong>t was placed out at 204 dd, 71 dd and 7 days later than the first. In 1991 the experiment<br />

was begun 10 calendar days earlier than in 1989 but at a phenologically later point, 210 dd, about the time of budbreak<br />

on the early flushing trees (Nienstaedt and King 1970). Hence, the first coh<strong>or</strong>t in 1991 was m<strong>or</strong>e nearly equivalent to the<br />

second coh<strong>or</strong>t in 1989. The second coh<strong>or</strong>t on fir in 1991 was placed on the host plants at 262 dd thereby having no 1989<br />

equivalents. Finally, the single coh<strong>or</strong>t on spruce in 1991 was placed on the trees at 280 degree days, about the time of<br />

budbreak f<strong>or</strong> the later flushing spruce trees (Nienstaedt and King 1970). In addition, in 1991, just after placing the first<br />

coh<strong>or</strong>t of second instars on the trees, the weather turned cool and wet keeping the young insects in their overwintering<br />

hibernaculum until nearly 1 week later when the next coh<strong>or</strong>t was being placed on the trees.<br />

Although <strong>this</strong> study attempted to examine the combined effects of staminate flowering by host phenology on the<br />

perf<strong>or</strong>mance of spruce budw<strong>or</strong>m, we were unable to execute the experiment in perfect phenological duplication that would<br />

have allowed the most powerful tests of the hypotheses.<br />

Flower by Phenology Effects<br />

, The data clearly suggest that the survival of budw<strong>or</strong>ms is dependent on the flowering x phenology interaction of its<br />

host plants. In the case of fir, abundant flowering probably enhances survival most significantly on the later flushing trees,<br />

i.e., late relative to budw<strong>or</strong>m emergence (they inevitably emerge bef<strong>or</strong>e their hosts break bud). This was apparent especially<br />

in 1989. In that year all flowering branches supp<strong>or</strong>ted higher survival than nonflowering branches regardless of tree phenology<br />

class, and regardless of coh<strong>or</strong>t timing. In 1991 there was no apparent flowering effect on survival except f<strong>or</strong> the late<br />

flushing trees, especially in the case of the second coh<strong>or</strong>t. Likewise, on spruce there was a strong tendency, though only<br />

nearly significant (p


There was practically no evidence to suggest that flowering enhances the growth and development rate of budw<strong>or</strong>ms,<br />

not on fir, n<strong>or</strong> on spruce. Likewise, there seems to be no strong evidence that there is any kind of F x P effect on growth<br />

processes. To the contrary, the data on spruce, though not quite statistically significant (p


FOLIVORE FEEDING ON MALE CONIFER FLOWERS °<br />

DEFENCE AVOIDANCE OR BET-HEDGING?<br />

THOMAS SECHER JENSEN<br />

Institute of Biological Sciences, Department of Zoology<br />

Aarhus University, DK-8000 Aarhus C, Denmark<br />

INTRODUCTION<br />

In the fifties and sixties a group of British researchers led by G.C.Varley made pioneering studies on the life hist<strong>or</strong>y<br />

of the winter moth (Varley and Gradwell 1970). Among many other things, they concluded that the key to understanding the<br />

main m<strong>or</strong>tality in <strong>this</strong> species was the timing between egg hatch and bud-burst. Larvae which hatch on trees, having buds<br />

firmly closed, fail to find food there. Instead such larvae disperse by spinning a thread of silk, and are blown away to an<br />

uncertain destiny. Larvae that hatch when buds have just burst, survive well. This first model was further developed by<br />

Feeny (1970) who in another pioneering study argued further that larvae which hatch after bud-burst face certain difficulties,<br />

<strong>this</strong> time in relation to nutritional conditions, mainly decreased nitrogen and water, and increased tannin concentrations.<br />

In summary these fact<strong>or</strong>s indicate that in Quercus robur, there is a "phenological window" represented by recently<br />

burst, highly nutritious and po<strong>or</strong>ly-defended leaves. In oak many defoliat<strong>or</strong>s, especially lepidopteran species, take advantage<br />

of <strong>this</strong> window and often reach high densities, even outbreak levels, at that time. Also species richness is highest in the early<br />

spring. However, <strong>this</strong> situation in Quercus robur does not seem to apply in a number of other tree species. The period of<br />

shoot growth in oak is rather sh<strong>or</strong>t, and Niemel/i and Haukioja (1982) demonstrated that in many other deciduous trees the<br />

period of shoot growth is much longer and seems to give defoliat<strong>or</strong>s a chance to find a much broader phenological window.<br />

In <strong>this</strong> paper the nutritional status of two coniferous genera, pine, Pinus, and spruce, Picea, are considered in relation<br />

to the phenological window, including the situation when alternative food, especially flowering buds are available. The<br />

insect species studied has mainly been the nun moth, Lymantria monacha, a close relative of the gypsy moth. The nun moth<br />

is a polyphagous species that often defoliates large areas of conifer f<strong>or</strong>est in Central Europe (Jensen 1991).<br />

METHODS<br />

Rearings and experiments of nun moth larvae and other insects were perf<strong>or</strong>med in a lab<strong>or</strong>at<strong>or</strong>y at constant 20°C.<br />

Eggs were collected in the f<strong>or</strong>est and brought to the lab<strong>or</strong>at<strong>or</strong>y during winter. In growth and survival experiments nun moths<br />

were kept in batches of ten in the first three instars; fourth and fifth instars were reared singly. Larvae were fed needles and<br />

flowers attached to excised twigs put in water.<br />

Chemical analysis were perf<strong>or</strong>med by means of micro-Kjeldahl analysis (nitrogen) and gas-chromatography (carbohydrates,<br />

cyclites and phenolic acids). F<strong>or</strong> details see Jensen (I988).<br />

RESULTS<br />

Spruce<br />

The genus Picea comprises about 30 species in the n<strong>or</strong>thern hemisphere. In parts of Europe, N<strong>or</strong>way Spruce, Picea<br />

abies, is the native species, but it is also widely planted, along with the introduced N<strong>or</strong>th American species Sitka spruce,<br />

Picea sitchensis. The buds burst rather early in spring along with a lot of deciduous trees, and the new shoots and needles are<br />

Mattson, W.J., Niemel_i, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

104


soft f<strong>or</strong> quite some time. It has 2-5 year classes of older needles. Nutritional content of the new needles is high: total N<br />

reaches 4.5% dwt, carbohydrate levels 250 nanomoles/g dwt (Jensen 1991). When needles get older, nitrogen levels drop to<br />

around 1%, whereas carbohydrate levels increase somewhat (Table 1).<br />

Table 1.--Concentrations of nutrients and secondary compounds in needles and male flowers of spruce, Picea abies.<br />

Total Amino Shikimic Quinic Pinitol Hexoses Inositol Cathecin Total<br />

Plant .part Nitrogen acids acid acid Phenolics<br />

% ................... n moles/mg dwt ...................<br />

Male flowers 5 _900 29 213 134 575 18 4<br />

Old top needles 1.2 _450 224 52 61 390 7 12 68<br />

New top needles 4 1,t20 108 625 133 248 17 3 35<br />

In Europe a substantial number of insect species are flush-feeders on spruce although some of them also consume<br />

older foliage, especially late in year <strong>or</strong> at outbreak densities (Table 2). This situation mainly applies to Lepidoptera although<br />

certain sawfly species also are flush-feeders. Another group feeds exclusively on older needles and is mainly comprised of<br />

Hymenoptera. In general, most spruce flush-feeders (e.g., L. monacha) perf<strong>or</strong>m substantially better when reared on the new,<br />

nutritious needles than on the old ones. Most newly hatched larvae die when put on old needles probably because their<br />

mandibles are not strong enough to cope with the hard needles (Jensen 1992). On the other hand, if third instar larvae are put<br />

on old needles, they survive and perf<strong>or</strong>m reasonably well. Preference indices f<strong>or</strong> new needles (Table 3) decrease with instar<br />

number, indicating that larger larvae do not require young needles.<br />

Table 2.--Needle age preferences of insect larvae feeding on N<strong>or</strong>way spruce, Picea abies.<br />

Needle age class<br />

Insect species Newly flushed Current Old<br />

Lepidoptera<br />

Parasyndemis histrionana * (*) _'<br />

Eana argentata * * (*)<br />

Epinotia tedella (*) *<br />

Epinotia nanana * * (*)<br />

Epinotia pygmaea *<br />

Zeiraphera ratzeburgiana *<br />

Orgya antiqua * * (*)<br />

Lymantria monacha * * (*)<br />

Hymenoptera<br />

Cephalcia abietis *<br />

Cephalcia arvensis * *<br />

Gilpinia hercyniae *<br />

Pristiph<strong>or</strong>a abietina *<br />

Pristiph<strong>or</strong>a ambigua *<br />

Pristiph<strong>or</strong>a saxeseni * (*)<br />

Pachynematus scutellatus * * (*)<br />

" * indicates unequivocal primary preferences, (*) indicates secondary preference, often by late instars.<br />

105


Table 3.--Lymantria monacha preference f<strong>or</strong> new needles of spruce, Picea abies.<br />

Instar Preference index n P<br />

I 1.00 + 0.00 20 < 0.001<br />

II 1.00 + 0.00 17 < 0.001<br />

III 0.84 + 0.32 9 < 0.020<br />

IV 0.63 + 0.15 21 < 0.020<br />

V 0.54 + 0.26 21 n.s.<br />

However, as indicated in Table 2, some species actually prefer the old needles, and to such an extent that they will die<br />

if fed the tender, nutritious new needles. An example of <strong>this</strong> is the sawfly, Gilpinia hercyniae, known in N<strong>or</strong>th America as<br />

the European Spruce Sawfly. It is a f<strong>or</strong>mer pest in Canada, but not in Europe. When offered only new foliage, <strong>this</strong> defoliato'r<br />

dies without eating (Jensen 1988). Experiments have shown that one possible deterring compound to G. hercyniae could be<br />

quinic acid, which together with shikimic acid is responsible f<strong>or</strong> much of the acidity of these new needles. Quinic acid is<br />

found in huge amounts in the new needles, but in low concentrations in the old ones. When quinic acid was added to old<br />

(preferred) needles in an experiment (Jensen 1988), G. hercyniae refused to eat after concentrations were increased to the<br />

levels of new needles.<br />

With respect to the phenological window, insects like G. hercyniae, feeding on old needles have no problems. In<br />

contrast, the flush-feeders must find the phenological window. If they hatch very early, even dispersal will not help if the<br />

food resource is not yet present. The only way to survive is to find an alternative food. In the dense monocultures of<br />

European spruce f<strong>or</strong>ests it is not easy to find other host species. However, occasionally, i.e., with 2-7 year intervals flowering<br />

buds might offer <strong>this</strong> alternative. The pollen buds are prime food, being soft, with high nutritional content and rather low<br />

content of secondary compounds (Table 1). Lab<strong>or</strong>at<strong>or</strong>y experiments on newly emerged L. monacha showed that these larvae<br />

are very fond of male flowers. When they hatch early and are reared on branches with only old needles, they die. If male<br />

infl<strong>or</strong>escences are present they survive and grow well (Fig. 1). Once started on flowers, they can eat new and old needles.<br />

Later in the season the larvae may struggle f<strong>or</strong> some time until vegetative buds burst.<br />

0.5 - Weight (g)<br />

0.3 - ""<br />

0.4 - o,,o<br />

0.2 - ,," __"<br />

0.1 - _o",, ,. "<br />

0.02<br />

/<br />

0.04 0.00<br />

_/_/ ,,"<br />

16. May<br />

o Male flowers ->New needles<br />

® Male flowers ->Old needles<br />

Old needles ->Old needles<br />

Figure 1.--Lymantria monacha growth on male flowers and needles of N<strong>or</strong>way spruce (Picea abies).<br />

106


Pine<br />

tn pine species the new needles emerge rather late in spring. Old needles are tough and of low nutritive value. This<br />

means that there is a long period with few feeding opp<strong>or</strong>tunities f<strong>or</strong> herbiv<strong>or</strong>es, except f<strong>or</strong> the existence of male infl<strong>or</strong>escences.<br />

In most places, pines set flower buds much m<strong>or</strong>e consistently than do spruce.<br />

New pine needles are less nutritive than spruce, nitrogen reaches only 2-2.5% of dwt. The toxins occurring in new<br />

needles are potent and it has been demonstrated (Ikeda et al. 1977) that they reduce growth rate, increase m<strong>or</strong>tality, and<br />

extend larval period. Not surprisingly, most defoliat<strong>or</strong> species feed on old needles <strong>or</strong> current year needles late in the year<br />

(Larsson and Tenow 1980, Table 4).<br />

Table 4.--Needle age preferences of insect larvae feeding on Scots pine, Pinus sylvestris.<br />

Insect species Needle age class<br />

Current Old<br />

Lepidoptera<br />

Bupalus pM[aria * *<br />

Cedestis spp *<br />

Cidaria firmata *<br />

Dendrolimus pini * *<br />

Ellopia fasciaria *<br />

Hyloicus pinastri *<br />

Panolis flammea * *<br />

Semio<strong>this</strong>a liturata *<br />

Lymantria monacha *<br />

Hymenoptera<br />

Diprion pini * *<br />

Diprion ximilis *<br />

Gilpinia frutet<strong>or</strong>um *<br />

Microdiprion pallipes * *<br />

Neodiprion sertifer *<br />

Nun moth larvae n<strong>or</strong>mally hatch pri<strong>or</strong> to bud-burst, leaving them with only old pine needles f<strong>or</strong> food. Experiments in<br />

field enclosures and in the lab<strong>or</strong>at<strong>or</strong>y with first instar larvae on whole branches <strong>or</strong> twigs, showed almost 100% m<strong>or</strong>tality.<br />

Only if the bagged branches included male flowers were larvae able to survive. In the lab<strong>or</strong>at<strong>or</strong>y, larvae survive and develop<br />

only if they are put on flowering branches in the first part of their life, irrespective of pine species (P cont<strong>or</strong>ta, P. sylvestris,<br />

P. mugo) (Table 5). Surviving larvae that later were fed new <strong>or</strong> old needles, perf<strong>or</strong>med equally well (Fig. 2).<br />

Table 5.--Percent survival of nun moth larvae on pines. T = 17 days, n = 85 per trial.<br />

P. silvestris P. mugo P. cont<strong>or</strong>ta<br />

Flowering 71.0 50.8 14.6<br />

Non-flowering 15.0 16.7 0.0<br />

107


500-<br />

400-<br />

i 300<br />

v<br />

¢<br />

m<br />

a 200-<br />

I00<br />

Lymantrlamonacha<br />

on Pinusmugo<br />

Nonflowering<br />

0 2O 40<br />

Days<br />

Flo_ering,old<br />

Figure 2:--,L_ymantria monacha growth on flowering and nonflowering branches of mountain pine (Pinus mugo). On<br />

flowering branches larvae were later separated into feeding groups on new and old needles.


DISCUSSION<br />

Based on the above findings it seems necessary to add another set of curves to the Varley-Feeny phenological<br />

window, curves that represent the opening of male flowering buds (Fig. 3). These curves are fundamentally different in the<br />

pine and spruce. In spruce, where male flowers burst very early in spring, it can be noted that flush feeders simply get a<br />

broader phenological window in years with male flowers. This means that larvae which hatch early in such a year get a<br />

larger probability of finding highly nutritious food. However a strategy of early hatching is a risky business, because chances<br />

are high that there will be no flowers. On the other hand, if they hatch late the risk is small, but the nutritional quality is<br />

lower and hence development time and predation risk will be higher.<br />

,,.__../<br />

4-_<br />

___c2<br />

Spr'uce<br />

,_ s#.._ - ,-.....<br />

o" egg bud<br />

......<br />

Lf)<br />

c_ Pi_e - ea_'l7 hatch<br />

M i<br />

egg o" bud<br />

,' /-X,"-"/ , \<br />

Cxb<br />

_ Pine -late hatch m<br />

r.D<br />

o c; bud egg<br />

U9<br />

" I,'-',<br />

/'""..../qk<br />

0<br />

q _ tlltlfllk<br />

,' /', /\ lk<br />

,' ..: ',,/ "..lllltllllllllllk<br />

, .., / ,, NWIIIIIIIIIIIN<br />

_ .... / % III ItIi"N,ItlIIIIII IIII1_<br />

,..."" ..... / _ IIITTtt,14LLIJIIIIIIII _<br />

Figure 3._Modification of the Varley-Gradwell curves f<strong>or</strong> the "phenological window": insect egg hatching, male flower and<br />

bud opening on spruce and pine.<br />

LID<br />

109


From the point of view of trees, the evolutionary response in terms of defense seems to be consistent with a concept<br />

k_own from deciduous species like oak <strong>or</strong> beech, as masting <strong>or</strong> mast seeding (Silvertown 1981). tn other w<strong>or</strong>ds, in some<br />

years there will be high beech-mast <strong>or</strong> ac<strong>or</strong>n production, but no seeds in the years in between. In the latter years, pest<br />

populations will decrease markedly. The strategy has probably evolved against seed predat<strong>or</strong>s but it could also w<strong>or</strong>k against<br />

_omedefoliat<strong>or</strong>s.<br />

The second line of defense is simply the variability of bud burst (Mitcherlich and Wellenstein 1942). Large variation<br />

in bud burst between provenances exist, but also large variation occurs within provenances. However, variability as a defense<br />

will mean that some larvae always will survive.<br />

The third line of defense against flush-feeders is a chemical one. Apparently, rather few spruce herbiv<strong>or</strong>es are<br />

deterred from feeding on newly burst needles to such an extent that they prefer old needles (Schopf 1986).<br />

In contrast, in pine it seems essential that flush feeders avoid plant defenses and here male flowers are of vitae<br />

imp<strong>or</strong>tance. In pine, male flowers are a m<strong>or</strong>e predictable resource than in spruce. Larvae that hatch after the flowering<br />

period will die and hence the risk of early hatch is less because male flowers are most often present. However, the strategy<br />

might fail; they might hatch on a tree without flowers, <strong>or</strong> w<strong>or</strong>se in a year without flowers in the whole stand.<br />

In the context of the phenological window it is actually necessary to consider the choices made by the female moth.<br />

Where does she deposit her eggs? Does she in fact spread the risk so that at least some of her larvae will survive? But at the<br />

same _ime,does she abandon an optimal strategy of maximizing sh<strong>or</strong>t term fitness? Essential to all these considerations are<br />

_hedevelopment time of buds and eggs, both fact<strong>or</strong>s being related to spring temperatures, although probably not in the same<br />

fashioa. I hypothesize that the position of the eggs is a key fact<strong>or</strong> in <strong>this</strong> synchronization simply because the temperature<br />

varies so much between microsites.<br />

1#mantria monacha females deposit their egg clusters mainly on the tree trunk under bark scales. Egg numbers and<br />

egg hatch differ consideraNy between n<strong>or</strong>th- and south-facing sides of the trunk (Raae 1979), and it might be that <strong>this</strong> is the<br />

mechanism the females employ when they spread the risk. This could be the way whereby females ensure that some of their<br />

progeny in <strong>this</strong> variable environment will hatch when the phenological window is open. Another bet-hedging mechanism<br />

couid be that after laying their first batch and becoming lighter, females fly to another tree and deposit the rest of their eggs.<br />

If the first tree is in the middle of a dense stand and the next is at the southern edge, the temperature regime is different and<br />

acc<strong>or</strong>dingly the bets change.<br />

litis interesting to observe that many flush feeders have wingless females <strong>or</strong> have reduced flight capability due to<br />

their heavy loads of eggs. In L monacha, females are winged but po<strong>or</strong> fliers and it is likely that the females lay their eggs<br />

close to if not directly on the tree where they pupated, thereby increasing the likelihood of synchronization between the egg<br />

i_atchand bud burst.<br />

In conclusion, avoiding defenses of new needles is absolutely necessary f<strong>or</strong> pine defoliat<strong>or</strong>s but of min<strong>or</strong> imp<strong>or</strong>tance<br />

_o spruce de_bliat<strong>or</strong>s. Bet hedging might prove to be the evolutionary response by the defoliat<strong>or</strong>s in <strong>or</strong>der to find the host's<br />

phenological window.<br />

LITERATURE CITED<br />

DEN BOER, RJ. 1968. Spreading of risk and stabilization of animal numbers. Acta Biothe<strong>or</strong>etica 18:165-194.<br />

FEEENY, P 1970. Seasonal changes in oak leaf tannins and nutrients as a cause of spring feeding by winter moth caterpillars.<br />

Ecology 51' 565-582.<br />

KEDA, T., MATSUMURA, E, and BENJAMIN, D.M. 1977. Chemical basis f<strong>or</strong> feeding adaptation of pine sawflies,<br />

,_e_:dlpr_on rug(fr<strong>or</strong>ls and Neodiprion swainei. Science 197: 497-498.<br />

JENSEN, T.S. 1988. Variability of N<strong>or</strong>way spruce needles; perf<strong>or</strong>mance of spruce sawflies (Gilpinia hercyniae). Oecologia<br />

77:313-32(t.


JENSEN, T.S. 199 la. Integrated pest management of the nun moth, Lymatltria monacha (Lepidoptera: Lymantriidae) in<br />

Denmark. F<strong>or</strong>. Ecol. Mgmt. 39: 29-34.<br />

JENSEN, T.S. t99 lb. Patterns of nutrient utilization in the needle feeding guild. In Baranchikov et at., eds. F<strong>or</strong>est insect<br />

guilds: Patterns of interaction with host trees. GTR-NE-153. Radn<strong>or</strong>. PA: U.S. Department of Agriculture, F<strong>or</strong>est<br />

Service.<br />

JENSEN, T.S. 1992. The role of plant stress in the population dynamics of nun moths. IUFRO Conference Zakopane,<br />

Poland 1991.<br />

LARSSON, S. and TENOW, O. 1980. Needle-eating insects and grazing dynamics in a mature Scots pine f<strong>or</strong>est in Central<br />

Sweden. In Persson, T., ed. Structure and function of n<strong>or</strong>thern coniferous f<strong>or</strong>ests: an ecosystem study. Ecol. Bull. 32:<br />

269-306.<br />

MITCHERLICH, H. and WELLENSTEIN, G. 1942. Die Nonne an Frtih- und Sp_ittreiberf<strong>or</strong>men der Fichte. Mon. Z. ang.<br />

Ent. 15: 94-125.<br />

NIEMEL,_, R and HAUKtOJA, E. 11982. Seasonal patterns in species richness of herbiv<strong>or</strong>es: Macrolepidopteran larvae on<br />

Finnish deciduous trees. Ecol. Entomol. 7: 169-175.<br />

PHILIPPI, T. and SEGER, J. 1989. Hedging one's evolutionary bets, revisited. TREE 4:41:44.<br />

RAAE, K. 1979. The distribution of nun moth eggs in old spruce stands. Unpubl. rep. Agricultural Univ., Copenhagen.<br />

SCHONHERR, J. 1989. Outbreak characteristics of Lymantriids. In Wallner, W.E. and McManus, K.A., ed. Lymantriidae:<br />

A comparison of features of new and old w<strong>or</strong>ld tussock moths. GTR-NE-123. Radn<strong>or</strong>, PA: U.S. Department of<br />

Agriculture, F<strong>or</strong>est Service.<br />

SCHOPE R. 1986. The effect of secondary needle compounds on the development of phytophagous insects. F<strong>or</strong>. Ecol.<br />

Man. 15: 54-64.<br />

SEGER, J. and BROCKMANN, H.J. 1987. What is bet-hedging? Oxf<strong>or</strong>d Surv. Evol. Biol. 4: 182-211.<br />

SILWERTOWN, J.W. 1981. The evolutionary ecology of mast seeding in trees. Biol. J. Linn. Soc. 14: 235-240.<br />

VARLEY, G.C. and GRADWELL, G.R. 1970. Recent advances in insect population dynamics. Ann. Rev. Ent. 15: 1-24.<br />

111


THE BLACKMARGINED APHID AS A KEYSTONE SPECIES:<br />

A PREDATOR ATTRACTOR REDRESSING NATURAL ENEMY<br />

IMBALANCES IN PECAN SYSTEMS<br />

M.K. HARRIS and T. LI<br />

Department of Entomology, Texas A&M University, College <strong>Station</strong>, Texas 77843, USA<br />

INTRODUCTION<br />

Plants defend themselves against herbiv<strong>or</strong>es in many ways. Conceptualization of these defenses has resulted in many<br />

terms and contexts to describe them (see Harris and Frederiksen 1984). Snelling (1941) listed 15 categ<strong>or</strong>ies. Painter (1951,<br />

Fig. 3) depicted 11 plant fact<strong>or</strong>s interacting with 8 insect fact<strong>or</strong>s mitigated by 8 environmental fact<strong>or</strong>s influencing 5 insectplant<br />

interaction fact<strong>or</strong>s as possible causes of resistance to insects, and then consolidated these fact<strong>or</strong>s into three resistance<br />

mechanisms: tolerance, antibiosis, and preference. These mechanisms were, by definition, heritable, effective in isolation,<br />

and particularly apt f<strong>or</strong> usage in agriculture where producing a crop of good quality in the presence of the insect was paramount.<br />

Harris (1980) proposed a succinct and natural characterization of plant defense, listing escape in space and time,<br />

accommodation, confrontation, and biological associations as the primary mechanisms to consider.<br />

Biological Associations<br />

Of these, biological associations are the most complex and difficult to demonstrate as natural defense mechanisms,<br />

because both the genetics of the plant and the arthropod are involved. Perhaps the best known case of a natural biological<br />

association providing plants with a defense against herbiv<strong>or</strong>es is the ant-acacias. Howe and Westley (1988) review <strong>this</strong> and<br />

other lesser known ant-plant associations, as well as other biological associations including those imp<strong>or</strong>tant in dispersal,<br />

pollination, etc. Biological associations mediated by the genetics of the plant clearly play an imp<strong>or</strong>tant role in nature, yet<br />

they have seldom been deliberately exploited to a significant degree in agriculture. Exceptions could be c<strong>or</strong>n leaf aphid,<br />

Rhopalosiphum maidis (Fitch) (Homoptera: Aphidae) on c<strong>or</strong>n and s<strong>or</strong>ghum, and apple rust mite Aculus schlechtendali<br />

(Nalepa) (Acarina: Eriophyidae) on apple, which are relatively innocuous at moderate densities and, if left alone, will often<br />

supp<strong>or</strong>t densities of natural enemies that also suppress m<strong>or</strong>e pestiferous species of aphids and mites, respectively (Flint and<br />

van den Bosch 1977). The agricultural imp<strong>or</strong>tance of c<strong>or</strong>n leaf aphid and apple rust mite is clear, but the <strong>or</strong>igin and role of<br />

these biological associations in natural systems are presently unknown. From an agricultural perspective, <strong>this</strong> area combines<br />

aspects of biological control and host plant resistancentwo subdisciplines of entomology that lack a common paradigm.<br />

BLACKMARGINED APHID AND PECAN SYSTEMS<br />

This paper proposes another example of a biological association, viz. how pecan, Carya illinoensis (Wang.) K. Koch<br />

(Juglandaceae), interacts with the blackmargined aphid, Monellia caryella (Fitch) (Homoptera: Aphidae), to influence other<br />

biological associations that may impact plant defense within the context of the matrices of interactions noted above.<br />

Pecan is a deciduous, monoecious, wind-pollinated, woody perennial that occupies alluvial soils from western Texas<br />

to the Mississippi Valley on the east and from southern Illinois into Mexico (Little 1971). Pecan trees can live m<strong>or</strong>e than 200<br />

years, grow to heights exceeding 35 m, and can constitute as much as 50% of the tree canopy in their natural habitat (Maggio<br />

et al. 1991). Reproduction is by seed (=200/kg) abundantly produced every 2-7 years (< 500 kg/ha) in natural populations,<br />

and each tree in nature is genetically distinct (Harris 1988). Leaves are compound and occur at a density of about 3 million<br />

leaves <strong>or</strong> roughly 50,000 m2/ha.<br />

Mattson, W.J., Niemel/i, E, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

112


Pecan domestication has been a gradual process, initially consisting of thinning riverine woodlands to pure stands of<br />

pre-existing pecan trees. Wild pecan nuts are commercially attractive and usually preferred by bakers and confectioners<br />

because they are typically priced about 25% less than selected varieties; they are about 40% smaller, allowing the whole<br />

halves to be pleasingly presented on small cakes, candies, etc.; and they are comparable, if not superi<strong>or</strong>, in taste to selected<br />

varieties. M<strong>or</strong>e than 60% of current nut production in Texas comes from such trees with the remainder produced by vegetatively<br />

propagated varieties. Thus, the pecan landscape throughout the indigenous range consists of wild trees growing in<br />

mixed species woodlands, native pecan that has been thinned from woodlands, and vegetatively propagated <strong>or</strong>chards often<br />

found adjacent to one another. Management programs like fertilization, pesticide application, irrigation, and pruning range<br />

from being most intensive in <strong>or</strong>chards to virtually nonexistent in mixed species woodlands. This variety of habitats provides<br />

unique opp<strong>or</strong>tunities to examine many questions including effects of plant domestication on arthropods. Most investigations<br />

of the pecan arthropod complex have been spurred by economic considerations and have concentrated on species that pose<br />

economic threats <strong>or</strong> economic relief (Harris 1983). However, given the long, close association of the arthropod complex with<br />

the pecan and the physical proximity of the wild and cultivated varieties, the opp<strong>or</strong>tunity to interpret results of pecan arthropod<br />

interactions from other perspectives invites attention. The arthropod complex consists of m<strong>or</strong>e than a 100 phytophagous,<br />

predat<strong>or</strong>, and parasite species that have been associated with pecan from the earliest of times (Harris 1983).<br />

The blackmargined aphid (BMA) is a monophagous, multivoltine, obligat<strong>or</strong>ily alate, phloem feeder that overwinters<br />

as eggs placed in bark crevices and parthenogenetically infests pecan from sh<strong>or</strong>tly after budbreak until the fall some 7-8<br />

months later, when males are produced, mating occurs, and the overwintering egg population is established (Harris 1983).<br />

Seasonal phenology on wild trees typically consists of BMA densities initially below 1/leaf increasing to between 1-10<br />

aphids/leaf f<strong>or</strong> a 2-3 week period during the summer, and then returning to densities below l/leaf f<strong>or</strong> the remainder of the<br />

season. Orchard trees typically experience _>rivefold higher densities and exhibit BMA population increases earlier in the<br />

season that last f<strong>or</strong> 3-4 weeks (Liao et al. 1984).<br />

BLACKMARGINED APHID AS A KEYSTONE SPECIES<br />

Effects of BMA on nut production appear to be negligible in wild trees (Liao and Harris I984), and with one exception<br />

(Wood et al. 1987), on <strong>or</strong>chard trees. However, routine use of insecticide in the early season typically results in epidemics<br />

of pecan aphids (PA) (Monelliopsispecanis Bissell), mites, and leafminers (Harris 1988, 1991). The pecan aphid complex<br />

consisting of BMA and PA has particularly been considered a primary threat to nut production (Dutcher and Htay 1985,<br />

Beshears 11988). Population densities of BMA are typically an <strong>or</strong>der of magnitude lower than those of PA in such epidemics<br />

and, if the epidemic proceeds to defoliation, BMA is virtually absent during the latter stages (Bumroongsook and Harris<br />

1992).<br />

BMA outbreaks routinely occur once per season in nature, <strong>or</strong> can also be induced once following the use of broad<br />

spectrum insecticides. Cage studies show that BMA outbreaks occur 2-3 weeks after introducing BMA into natural enemy<br />

exclusion cages throughout the season on previously unexposed foliage (Edelson 1982, Liao and Harris 1984, Liao et al.<br />

1984, Liao et al. 1985, Bumroongsook and Harris 1992). These outbreaks naturally subside after 3-4 weeks, leaving intact<br />

photosynthetically active foliage (Bumroongsook and Harris 1992). BMA populations cannot reinfest the foliage f<strong>or</strong> at least<br />

a month after the first induced outbreak (Liao et al. 1984). BMA outbreaks can be curtailed by opening the exclusion cages<br />

bef<strong>or</strong>e the infestation has run its course <strong>or</strong> by introducing spiders <strong>or</strong> lady beetles into the cage at densities of about 1 per 10<br />

leaves (Liao et al. 1984). Lacewing, Chrysoperla rufilabris (Burmeister), oviposition increases as BMA densities increase.<br />

Lacewing egg densities are also higher on foliage exposed by opening cages containing an incipient outbreak of BMA (Liao<br />

et al. 1984). Spider densities increase as BMA densities increase, and spiders readily accept BMA as prey (Bumroongsook et<br />

al. 1992). Spider densities are also higher on foliage exposed by opening cages containing an incipient outbreak of BMA<br />

(Liao et al. 1984). M. pecanis is slower to outbreak on mature bearing trees (Bumroongsook and Harris 1992), and high<br />

densities result in defoliation whether the leaves have been conditioned by BMA <strong>or</strong> not, indicating these aphids are capable of<br />

causing m<strong>or</strong>e loss of foliage than BMA.<br />

These studies indicate that BMA may mitigate the impact of disruptions <strong>or</strong> natural declines of natural enemies in the<br />

pecan ecosystem by outbreaking and then attracting and reestablishing natural enemies to the system bef<strong>or</strong>e m<strong>or</strong>e insidious<br />

phytophages increase in density (Bumroongsook and Harris 1992). The effects of such a role would not be limited to the<br />

pecan aphid complex, but would extend to all phytophagous species serving as prey to the polyphagous lacewings, lady<br />

beetles, spiders, etc. (Liao et al. 1984), that respond to M. caryella.<br />

113


Wood et al. (1987) rep<strong>or</strong>t that M. caryella second and third instar nymphs excrete some f<strong>or</strong>tyfold m<strong>or</strong>e honeydew<br />

than other pecan aphids (other life stage comparisons are three- to tenfold) and noted that some 95% of the 301 joules (J) of<br />

energy removed from pecan by one M. caryella aphid during its lifetime was due to honeydew excretion. M. cao'el/a energy<br />

consumption exceeds by 20% the largest consumption found in a review by Llewellyn (1987) f<strong>or</strong> non-pecan aphids, and also<br />

has the lowest energy assimilation compared to other aphids. Honeydew is attractive to hundreds of species of insects<br />

(Klingauf 1987), including many predat<strong>or</strong>s and parasites. Tedders (1991) concluded pecan aphid honeydew "may be<br />

necessary f<strong>or</strong> attracting and retaining large numbers of many beneficial species." Honeydew also improved the surrounding<br />

substrate f<strong>or</strong> decomposers (Dixon 1985), enhancing the recycling of nutrients. These benefits must be balanced against the<br />

costs of M. cao'ella to pecan.<br />

Initial characterization of the costs of M. caryella requires accounting f<strong>or</strong> energy lost to the aphid in relation to the<br />

overall photosynthate production capacity of the pecan. Wood et al. (1987) provide the f<strong>or</strong>mer at 301.41 J f<strong>or</strong> one M.<br />

caryella over a 19.3 d life span and calculate 11.42 J/d are removed by the average aphid in a population. Anderson (1991)<br />

rep<strong>or</strong>ted pecan assimilates CO2 at a rate of 15.9 btmol/m2/s [Wood and Tedders (1986) rep<strong>or</strong>t 11 g tool COJmZ/s]. Annual<br />

photosynthate productivity (PP)/ha, theref<strong>or</strong>e, can be estimated by:<br />

Annual PP/ha = 2.65 bttool of glucose x 50,000 m2pecan foliage/ha x 6.25 x 106seconds/growing season x<br />

0.5 x 0.65 [(c<strong>or</strong>rected f<strong>or</strong> photosynthetic efficiency due to shading (0.5) and respiration (0.65)].<br />

Where CO 2 has been converted to glucose (1 glucose = 6 COz), pecan foliage/ha was averaged from Wood et<br />

al. (1987) who rep<strong>or</strong>ted 35,900 mZ/ha, Cutler (1976) who rep<strong>or</strong>ted 52,580 m2/ha and Lozano (1982) who<br />

found 62,800 m2/ha; a growing season was defined as 217 d with 8 hr of sunlight each day; photosynthetic<br />

efficiency was estimated at 50% and respiration costs at 35%.<br />

This results in a PP/ha of 2.69 x 10s mols of glucose annually <strong>or</strong> 185.98 x 106 kcal/ha (1 tool of glucose = 691 kcal).<br />

Pecan foliage routinely is exposed to BMA at a level of a 100 to 500 <strong>or</strong> so aphid days per compound leaf each season<br />

(Tedders 1978, Fl<strong>or</strong>es 1981, Li 1990, Liao and Harris 1984, Tedders and Wood 1985, Bumroongsook 1986, Mansour and<br />

Harris 1988). Pecan leaf density estimates vary from 2.2 to 3.6 x l0 bleaves/ha (Cutler 1976, Lozano 1982, Wood et al.<br />

1987). Using an average of 3.0 x 106leaves/ha and 500 M. caryella aphid days/leaf/season results in an energy cost of 17.12<br />

x 10'_J/ha annually (I 1.42 J/aphid day x 3.0 x 106leaves/ha x 500 aphid days). Since 1 J = 0.24 cal, <strong>this</strong> equates to 4.1 x 10 6<br />

kcal as an average cost of BMA on each ha of pecan each season.<br />

Theref<strong>or</strong>e, a 500 aphid day/leaf infestation of BMA each season would occur at a direct cost of about 2% (4.1 x 106<br />

+185.98 x 1106)of the photosynthate production of the pecan. This cost would be spread across all physiological needs of the<br />

plant with only a p<strong>or</strong>tion appearing as yield loss. Secondary effects of fact<strong>or</strong>s like saprophagous sooty mold growth and<br />

additional reductions in photosynthetic efficiency are not discussed because previous studies (Tedders and Smith 1976, Wood<br />

and Tedders 1986) indicate BMA densities of 500 aphid days/leatTseason are too low to cause measurable reductions via these<br />

secondary effects in mature bearing trees.<br />

This view of BMA is m<strong>or</strong>e sanguine than that presented by Wood et al. (1987). Resistance to pesticides in the pecan<br />

pest complex makes the prospect of maintaining BMA below its natural equilibrium level fraught with additional pesticide<br />

costs and other secondary pest problems like mites, leafminers, and other aphids (Harris 1991). Wood et al. (1987) rep<strong>or</strong>t an<br />

average of 10 aphicide sprays ($500/ha at $50/spray)required to achieve <strong>this</strong>, and Dutcher (1991) rep<strong>or</strong>ts 50 kg/ha rates of<br />

aldicarb ($330/ha at $6.60/kg) needed to achieve the same end f<strong>or</strong> aphid control from mid-July onward. Although these<br />

approaches may protect the = 2% of photosynthate production at risk from BMA, that loss represents 20 kg from an estimated<br />

average yield of 1,000 kg/ha if the entire BMA energy removed resulted in nut loss. Most insecticides currently used in<br />

pecans are highly toxic to most natural enemies of aphids (Mizell 1991) and in danger of being rendered ineffective due to<br />

aphid resistance to them (Harris 1991). We believe that the pecan aphid complex rarely poses a direct threat to pecan<br />

production unless disrupted by pesticides, especially in the early season and that the judicious use of pesticides f<strong>or</strong> key pests<br />

can usually avoid these disruptions. When the problem is "viewed from <strong>this</strong> perspective, pest managers may wish to reevaluate<br />

the risks posed by BMA compared to the benefits provided by sustaining natural enemies and enhancing decomposition.<br />

Further study is needed to determine whether a natural role of BMA in the pecan ecosystem is to re-establish waning<br />

densities of natural enemies.<br />

114


HARRIS, M.K. 1980. Arthropod-plant interactions related to agriculture, emphasizing host plant resistance, p. 23-51. In<br />

Biology and Breeding f<strong>or</strong> Resistance. Dept. of Agric. Comm. TAES, Texas A&M University. MP-1451.<br />

HARRIS, M.K. 1983. Integrated pest management of pecans. Ann. Rev. Entomol. 28:291-318.<br />

HARRIS, M.K. and FREDERIKSEN, R.A. 1984. Concepts and methods regarding host plant resistance to avthropods and<br />

pathogens. Ann. Rev. Phytopath. 22: 247-272.<br />

HARRIS, M.K. 1988. Pecan domestication and pecan arthropods, p. 207-225. In The Entomology of Indigenous and<br />

Naturalized Systems in Agriculture. Westview Press Inc., Boulder, CO.<br />

HARRIS, M.K. 1991. Pecan arthropod management,, p. 6-15. In Wood, B.W. and Payne, J.A., eds. Pecan Husbandry:<br />

Challenges and Opp<strong>or</strong>tunities. First National Pecan W<strong>or</strong>kshop Proceedings. ARS-96. U.S. Deptartment of Agriculture,<br />

Agriculture <strong>Research</strong> Service.<br />

HOWE, H.E, and WESTLEY, L.C. 1988. Ecological relationships of plants and animals. Oxf<strong>or</strong>d Univ. Press, Oxf<strong>or</strong>d.<br />

273 p.<br />

KLINGAUF, EA. 1987. Feeding, adaption and excretion, p. 225-253. In Minks, K.A. and Harrewign, E, eds. Aphids: Their<br />

Biology, Natural Enemies, and Control. Elsevier, New Y<strong>or</strong>k.<br />

LI, T. 1990. Population growth and fact<strong>or</strong>s affecting the seasonal abundance Monellia caryella (Fitch) and Monelliopsis<br />

pecanis Bissell in Central Texas. Ph.D. Dissertation, Texas A&M University, College <strong>Station</strong>, TX. 181 p.<br />

LIAO, H. and HARRIS, M.K. 1984. Population growth of the blackmargined aphid on pecan in the field. Agric. Ecol.<br />

Environ. 12: 353-361.<br />

LIAO, H., HARRIS, M.K., GILSTRAP, EE., DEAN, D.A., AGNEW, C.W., MICHELS, G.J., and MANSOUR, F. 1984.<br />

Natural enemies and other fact<strong>or</strong>s affecting seasonal abundance of the blackmargined aphid on pecan. Southw.<br />

Entomol. 9: 404-420.<br />

LIAO, H., HARRIS, M.K., GILSTRAP, EE., and MANSOUR, E 1985. Impact of natural enemies on the blackmargined<br />

pecan aphid, Monellia caryella (Homoptera: Aphidae). Environ. Entomol. 14: 122-126.<br />

LITTLE, E.L. 1971. Atlas of United States trees. Vol. 1. Conifers and imp<strong>or</strong>tant hardwoods. Misc. Publ. 1146. Washington,<br />

DC: U.S. Department of Agriculture, F<strong>or</strong>est Service.<br />

LLEWELLYN, M. 1987. Aphid energy budgets, p. 109-117. In Minks, A.K. and Harrewign, E, eds. Aphids: Their Biology,<br />

Natural Enemies, and Control Elsevier, New Y<strong>or</strong>k.<br />

LOZANO, R. 1982. A three-dimensional characterization of a bearing pecan tree. M.S. Thesis, H<strong>or</strong>ticulture, Texas A&M<br />

University, College <strong>Station</strong>, TX. 71 p.<br />

MAGGIO, R.C., HARRIS, M.K., INGLE, S.J., and DAVIS, M.R. 1991. A summary of the location, abundance, distribution,<br />

and condition of Carya on the Brazos and Col<strong>or</strong>ado River systems of Texas. Dept. of Agric. Comm. TAES, Texas<br />

A&M Univ. MP-1703.<br />

MANSOUR, E and HARRIS, M.K. 1988. Biology and phenology of the blackmargined aphid, Monellia caryella (Fitch), a<br />

new pest of pecan in Israel. Southw. Entomol. 13: 19-29.<br />

116


MIZELL, R.E 1991. Pesticides and beneficial insects: application of current knowledge and future needs, p. 47-54. In<br />

Wood, B.W. and Payne, J.A., eds. Pecan Husbandry: Challenges and Opp<strong>or</strong>tunities. First National Pecan W<strong>or</strong>kshop<br />

Proceedings. ARS-96. U.S. Deptartment of Agriculture, Agriculture <strong>Research</strong> Service.<br />

PAINTER, R.H. 1951. Insect resistance in crop plants. Univ. Press of Kansas. Lawrence, KS. 520 p.<br />

SNELLING, R.O. 1941. Resistance of plants to insect attack. Bot. Rev. 7: 543-586.<br />

TEDDERS, W.L. and SMITH, J.W. 1976. Shading effect on pecan by sooty mold growth. J. Econ. Entomol. 69: 551-553.<br />

TEDDERS, W.L. 1978. Imp<strong>or</strong>tant biological and m<strong>or</strong>phological characteristics of foliar feeding aphids of pecan. Tech. Bull.<br />

1529. Washington, DC: U.S. Department of Agriculture. 29 p.<br />

TEDDERS, W.L. and WOOD, B.W. 1985. Estimate of the influence of feeding by Monelliopsis pecanis and Monellia<br />

caryella (Homoptera: Aphididae) on the fruit, foliage, carbohydrate reserves, and tree productivity of mature 'Stuart'<br />

pecans. J. Econ. Entomol. 78: 642-646.<br />

TEDDERS, W.L. 1991. Alternative controls f<strong>or</strong> pecan insects, p. 77-83. In Pecan Husbandry. ARS-96. U.S. Deptartment of<br />

Agriculture, Agriculture <strong>Research</strong> Service.<br />

WOOD, B.W. and TEDDERS, W.L. 1986. Reduced net photosynthesis of leaves from mature pecan trees by three species<br />

of pecan aphids. J. Entomol. Sci. 21: 355-360.<br />

WOOD, B.W., TEDDERS, W.L., and DUTCHER, J.D. 1987. Energy drain by three pecan aphid species (Homoptera:<br />

Aphididae) and their influence on in-shell pecan production. Environ. Entomol. 16(5): 1045-1056.<br />

117


PONDEROSA PINE RESPONSE TO NITROGEN FERTILIZATION AND<br />

DEFOLIATION BY THE PANDORA MOTH_<br />

COLORADIA PANDORA BLAKE<br />

B.E. WICKMAN I, R.R. MASON z, and H.G. PAUL 2<br />

_Silviculture Lab<strong>or</strong>at<strong>or</strong>y, 1027 NW Trenton Avenue, Bend, OR 97701, USA<br />

2F<strong>or</strong>estry and Range Sciences Lab<strong>or</strong>at<strong>or</strong>y, 1401 Gekeler Lane, La Grande, OR 97850, USA<br />

INTRODUCTION<br />

The pand<strong>or</strong>a moth, Col<strong>or</strong>adia pand<strong>or</strong>a Blake (Lepidoptera: Saturniidae), is a well-known outbreak species on<br />

ponderosa, Pinus ponderosa Laws, and lodgepole pines, Pinus cont<strong>or</strong>ta Dougl., in the western United States (Mattson et al.<br />

1991). F<strong>or</strong> example, it has periodically caused severe defoliation in the pineries of central Oregon. Hist<strong>or</strong>ically, such<br />

outbreaks have been sh<strong>or</strong>t-lived and caused little impact. But, in the 1920's, an outbreak in the Klamath Falls region did<br />

significant damage to the old-growth ponderosa pines so that they became highly susceptible to bark beetle attacks (Patterson<br />

1929).<br />

The insect has a 2-year life cycle (Carolin and Knopf 1968). Moths flights n<strong>or</strong>mally occur in June and July of evennumbered<br />

years (those near Bend, OR occur in odd-numbered years)(Patterson 1929, Carolin and Knopf 1968, Schmid et al.<br />

1982). Eggs are deposited in clusters on both needles and bark. Hatch takes place in August and the young larvae feed<br />

gregariously in small colonies on current-year needles until limited by cold weather in late fall. Feeding continues on warm<br />

winter days on the south side of trees and on trees highly exposed to sunshine. Larvae eventually disperse and feed solitari|y.<br />

Feeding intensity increases sharply with the onset of warm spring weather, and defoliation eventually becomes noticeable in<br />

early June as the caterpillars grow larger. By early July, fully grown larvae drop to the f<strong>or</strong>est flo<strong>or</strong> and pupate in the upper<br />

several centimeters of soil where they spend nearly a full year.<br />

The current outbreak in Central Oregon (first seen in 1988) has spread south of as well as n<strong>or</strong>th to the city of Bend<br />

because of massive moth flights in 1991 and 1993. In May 1992, we collected larvae in the <strong>or</strong>iginal epicenter that were<br />

infected with a polyhedrosus virus. This usually p<strong>or</strong>tends the collapse of an outbreak, so 1994 could have been the year of a<br />

widespread virus epizootic.<br />

Because pand<strong>or</strong>a moth outbreaks occur only every 20-30 years, the present infestation aff<strong>or</strong>ded an opp<strong>or</strong>tunity to test<br />

the effects of fertilization on the trees and the insects in the nutrient deficient soils in the central Oregon pineries. We<br />

hypothesized that increasing nutrient availability might improve canopy growth, and net photosynthesis and thereby lead to<br />

elevated carbon-based defenses against herbiv<strong>or</strong>y. Increased tree growth might also offset the effects of defoliation.<br />

In another f<strong>or</strong>est-insect system, we found that fertilization may temp<strong>or</strong>arily reduce the effects of defoliation (Mason<br />

et al. 1992, Wickman et al. 1992). During an outbreak of the western spruce budw<strong>or</strong>m, Ch<strong>or</strong>istoneura occidentalis Freeman,<br />

both trees and insects were enhanced by fertilization, but trees m<strong>or</strong>e than budw<strong>or</strong>ms because they produced m<strong>or</strong>e new foliage<br />

than the budw<strong>or</strong>m could eat. Consequently, fertilized stands suffered less growth impact from defoliation than did untreated<br />

stands. We tested <strong>this</strong> conclusion further in the ponderosa pine-pand<strong>or</strong>a moth system.<br />

The exact test was to determine if a single treatment with nitrogen in the f<strong>or</strong>m of urea would significantly reduce the<br />

impact of pand<strong>or</strong>a moth defoliation in a thinned second-growth ponderosa pine stand. A secondary objective was to determine<br />

the effect of the treatment on growth and feeding behavi<strong>or</strong> of pand<strong>or</strong>a moth larvae and evaluate the chemical composition<br />

of foliage and larval frass.<br />

Mattson, W.J., Niemel_i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

118


METHODS<br />

Study Site<br />

The study took place in an active outbreak of the pand<strong>or</strong>a moth in the Deschutes National F<strong>or</strong>est about 20 km south<br />

of Bend, Oregon. The stand is a ponderosa pine/bitterbrush-manzanitJfescue plant community (Volland 1976). Lodgepole<br />

pine is the only other tree species present and is scarce. The <strong>or</strong>iginal ponderosa pine was logged in the 1930's and the second<br />

growth stand thinned in the 1970's. The average age of the trees at the time of our study was about 55 years and spacing<br />

averaged 3.6 by 3.6 m, but the latter differed across the area (the study area averaged 165 trees per acre).<br />

Elevation of the study area is 1,250 in, slope varies from 0 to 5 percent, and aspect is slightly south. The soil mantle<br />

is 0.9-m-deep pumice, which resulted from the Mount Mazama volcanic eruption about 6,700 years ago. Bare pumice is<br />

common and there is only a thin <strong>or</strong>ganic layer under the trees. Soils are particularly deficient in nitrogen, phosph<strong>or</strong>us, and<br />

sulfur (Cochran 1978). Annual precipitation is approximately 51 cm and occurs mostly as snowfall from October through<br />

March. Precipitation at Bend was 23 cm in the 1988-1989 season, and first snowfall occurred within a week after we applied<br />

fertilizer.<br />

Experimental Design and Treatments<br />

In fall 1988, 10pairs of circular 1/10th of an acre plots (11.35-m radius) were located in a 4 by 5 grid with plot<br />

centers approximately 40 m apart (Fig. 1). All trees on the plots were marked with metal tags. On average, treated plots<br />

experiment near Sun River<br />

on the Deschutes National<br />

Figure l.--Schematic F<strong>or</strong>est. F = fertilized layout of and C<br />

= control.<br />

_ _ @ _<br />

119


contained 17 trees and control plots 16 trees. The average tree diameter at breast height (dbh) was 29.5 cm. One plot in each<br />

pair was randomly selected f<strong>or</strong> fertilization. Analysis of results was on a paired plot basis. In October, 27.2 kg of urea (47%<br />

N) was applied by hand to the fertilized plot in each of the 10 pairs. This simulated an application rate of 350 kg N ha -_.<br />

Insect and Tree Measurements<br />

Effects of the treatment were evaluated by measurements of tree buds, foliage, radial increment, and insect biomass<br />

on treated and untreated plots. Samples f<strong>or</strong> foliage measurements and insect weights were collected periodically from tree<br />

crowns with pole pruners acc<strong>or</strong>ding to methods recommended by Schmid et al. (1982). Radial increment was evaluated at<br />

the end of the study from increment c<strong>or</strong>es taken from two trees per plot. All sample trees were selected from the center of<br />

plots to minimize edge effect. The schedule f<strong>or</strong> measurements and collections follows:<br />

Shoot and Insect Insect Radial<br />

Date foliage larvae pupae increment<br />

Fall 1989 X<br />

May 1990 X<br />

Fall 1990 X X<br />

Fall 1991 X<br />

May 1992 X<br />

Fall 1992 X X<br />

Larvae were sampled only once in 1990, using methods developed by Schmid et al. (1982). Sampling also was attempted in<br />

1992, but only a few larvae were found, indicating a virtual population collapse in the study area. Larvae were collected f<strong>or</strong><br />

weighing from 2 midcrown, 45-cm branch samples on each of 3 trees per plot. Larvae were preserved in 70% ethanol and<br />

oven dried at 45 °C to a constant weight. Mean dry weight per individual larva was calculated by dividing dry mass by<br />

density. Mean fresh weights of pupae were determined from a collection made from the soil of approximately 5 pupae per<br />

plot.<br />

Branch samples were collected each August <strong>or</strong> September from 1989 to 1992. Two 45-cm mid-crown branches per<br />

tree, from 3 trees per plot were cut:with pole pruners. In the lab<strong>or</strong>at<strong>or</strong>y, new buds were excised and current needles were<br />

stripped from the branches. These were oven dried at 45°C to a constant weight and expressed as foliage dry weight per<br />

centimeter of twig length. Mean foliage weight was summarized by plot and year. Buds were represented as weight per bud.<br />

In August 1992, at the end of the growing season, 2 c<strong>or</strong>es per tree were collected at dbh on 2 trees per plot to measure radial<br />

increment. C<strong>or</strong>es were mounted, sanded, and measured to the nearest 0.01 mm on an incremental measuring instrument<br />

interfaced with a desk top computer, as described in Wickman et al. (1992).<br />

Insect frass was collected from canvas panels placed under trees on 5 treated and 5 untreated plots at the peak of<br />

larval feeding in June 1990. Frass was air dried and analyzed f<strong>or</strong> total and available levels of N, P, K, and S at the University<br />

of Arizona, Tucson, in 1993. Dried foliage from 1989 and 1990 samples also was analyzed f<strong>or</strong> N, P, K, and S and available<br />

nutrients at the same lab<strong>or</strong>at<strong>or</strong>y in 1993. Dried 1989 needles were analyzed f<strong>or</strong> total reducing sugars and phenolics by Mr.<br />

B.A. Birr, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong>, East Lansing, MI, in 1994.<br />

Statistical Analysis<br />

Data were analyzed using paired t-tests on the randomly selected paired plots f<strong>or</strong> each year of measurement. Differences<br />

were judged to be significant when the probability was


the samples. Fertilization had substantial negative effects. Mean dry weights of individual larvae from fertilized plots were<br />

126.0 mg and 248.7 mg f<strong>or</strong> 4th and 5th instars respectively, compared to 196.5 mg and 360.3 mg, respectively, tbr untreated<br />

plots. These were 29.6% (p = 0.036) and 32.9% (p = 0.006) reductions, respectively, due to fertilization. Mean fresh weights<br />

of treated and untreated pupae in September, 1990 were 290.1 mg and 30l .4 rag, respectively, but were not significantly<br />

different (p = 0.56).<br />

Foliage and Bud Weights<br />

Foliage weight was consistently greater on the control plots from 1989 through 1992, but <strong>this</strong> difference was significant<br />

(p = 0.03) only in 1990 (Fig. 2). The declining weights of all foliage from 1989 through 1992 was probably due to the<br />

combined effects of chronic drought and repeated de|bliation. Visual defoliation ratings made in fall, 1990, were similar f<strong>or</strong><br />

both fertilized and control trees. The trend of bud weight was similar to that found f<strong>or</strong> foliage except in 1989 fertilized buds<br />

were heavier than the controls (Fig. 3). Buds from control plots were significantly heavier in 1991 and 1992 than fertilized<br />

buds (132.3 vs 120.8 mg, p = 0.053 and 69.5 vs 49.3 rag, p = 0.021, respectively). Again, the much smaller buds in 1992<br />

probably reflect combined effects of 2 defoliation episodes and 5 consecutive years of subn<strong>or</strong>mal precipitation.<br />

60<br />

50<br />

"*" 40<br />

..c: ---7<br />

._<br />

30<br />

(D<br />

03 : . . •<br />

._ 20<br />

0<br />

LL :<br />

1o / /<br />

0 -/ /<br />

1989 1990 1991 1992<br />

Year<br />

Figure 2.--Foliage per centimeter of shoot length. Each bar is an average of 10 plot means.<br />

Foliage and Frass Chemical Analyses<br />

Foliage from fertilized trees contained m<strong>or</strong>e total N in both 1989 and 1990 and m<strong>or</strong>e available N in 1989 (Table 1).<br />

Statistical analysis was not possible because the samples were pooled. Other nutrients (P, S, and K), apparently were similar<br />

in both fertilized and control foliage f<strong>or</strong> both years (Table 1).<br />

Larval frass collected in June 1990 was also analyzed f<strong>or</strong> nutrients (Table 2). Available N was over 6 times higher in<br />

frass from larvae feeding on fertilized trees. Total P and K in frass from fertilized trees was approximately half that from<br />

unfertilized trees. Because the samples were pooled, a statistical analysis was not possible, but these differences were<br />

striking.<br />

121


Figure 3.--Average weight per bud. Each bar is an average of 10 plot means.<br />

Table l.--Percent foliar nutrients in 1989 and 1990 needles (sample pooled from 10 fertilized plots and 10 control plots).<br />

122<br />

Total<br />

N P S K (ppm)<br />

1989 7,558<br />

Plots 1-10 (fertilized) 2.095 0.114 0.071<br />

11-20 (control) 1.568 0.127 0.075 7,471<br />

1990 8,641<br />

Plots 1-10 (fertilized) 2.012 0.t67 0.082<br />

11-20 (control) 1.750 0.185 0.088 9,995<br />

Available<br />

1989<br />

Plots l-t0 (fertilized) 0.0216 0.0690 0.0088 6,165<br />

11-20 (control) 0.0134 0.0581 0.0064 5,635<br />

1990<br />

Plots 1-10 (fertilized) 0.0206 0.1007 0.0039 7,082<br />

11-20 (control) 0.0216 0.0994 0.0052 7,739


Table 2.--Percent nutrient content of larval frass in 1990 (sample pooled from 5 fertilized plots and 5 control plots).<br />

Total Available<br />

N P S ppmK ppm N<br />

Fertilized 0.826 0.065 0.039 1,430 625<br />

Control 0.595 0.131 0.042 2,670 94<br />

There was no significant difference in the total reducing sugars (p = 0.36), <strong>or</strong> phenolics (p = 0.48) from fertilized<br />

trees and controls in 1989 foliage.<br />

Radial Growth<br />

Radial growth showed a steadily declining trend on both fertilized and control plots from 1988 through 1992, (Fig. 4).<br />

Fertilized trees, even though they produced less foliage and smaller buds during <strong>this</strong> period, had significantly (p


Fertilized and control trees all grew at essentially the same rates during the 16-year period pri<strong>or</strong> to fertilization (Fig.<br />

5). The downward trend starting in 1986, probably induced by precipitation deficits, was sharply reversed in 1990 f<strong>or</strong><br />

fertilized trees. Those trees eventually resumed their downward trend, but the spurt of growth in 1990 gave them a noticeable<br />

advantage over the untreated controls.<br />

3<br />

2.5<br />

Control<br />

--Fertilized .<br />

_ ,, \<br />

Treatment<br />

Application<br />

/ /" \<br />

E / X...<br />

t -.<br />

E I "..<br />

\ //\<br />

0 \ _ s \\" \<br />

,-- 1.5 ,'-<br />

(_ _ / " [] Precip.departures frommean \\<br />

(b<br />

+10 cm<br />

1t,,..,<br />

09<br />

0<br />

<<br />

cu 1<br />

0.5<br />

'\\<br />

-10cm<br />

,,o_,,o_,,,o_,,,o_,,,o_ ,o5,,o_,,,c_,,o5,o5_<br />

0 -----T ] l l T I I I [ t I I I i i T ] T I<br />

1972 1976 1980 1984 1988 1992<br />

Year<br />

Figure 5._Trends of precipitation and radial growth of ponderosa pine at dbh. Plotted points are averages of 10 plot means.<br />

DISCUSSION<br />

All stages of the pand<strong>or</strong>a moth are large and easily seen, and tree defoliation is spectacular during an outbreak. Such<br />

episodes, however, occur at intervals of 20 <strong>or</strong> 30 years, and tree m<strong>or</strong>tality as a result of feeding is usually negligible. Even<br />

though the pand<strong>or</strong>a moth makes an interesting test insect f<strong>or</strong> herbiv<strong>or</strong>e studies, its irregular, boom <strong>or</strong> bust dynamics limit its<br />

consistent availability. Because of life hist<strong>or</strong>y differences between western spruce budw<strong>or</strong>m and pand<strong>or</strong>a moth, comparisons<br />

of fertilizer effects are tenuous. The pand<strong>or</strong>a moth has a 2-year life cycle; theref<strong>or</strong>e, larvae, on alternate years, feed mostly<br />

on older needles rather than emerging new foliage so that the tree has a constant supply of new foliage. Conversely, the<br />

budw<strong>or</strong>m has a l=year life cycle and feeds on new needles annually and can repeatedly remove new foliage f<strong>or</strong> 7 to 10 years.<br />

We found that larvae responded negatively in 1990, at least in terms of individual weights, to the fertilizer treatment.<br />

This occurred even though available N measured in 1989 foliage was almost 50% higher in fertilized plots. The opposite<br />

effects on larval weights were rep<strong>or</strong>ted f<strong>or</strong> a similar study of western spruce budw<strong>or</strong>m in Oregon (Mason et al. 1992, Waring<br />

et aL 1992). Available N in 1990 insect frass was over 6 times higher on fertilized plots. A possible explanation is that<br />

pand<strong>or</strong>a moth larvae may be differentially abs<strong>or</strong>bing proteins <strong>or</strong> amino acids from fertilized compared to unfertilized foliage<br />

and concentrating N in the frass. Why the increased nutrient resource did not result in larger individuals is an interesting<br />

question. Perhaps nutrients were imbalanced.<br />

124


Some tree responses to fertilization were also different than in other recent studies. Additional N provided by<br />

fertilization was expected to increase foliage production based on results of a study of fir infested with western spruce<br />

budw<strong>or</strong>m (Wickman et al. 1992). We found fertilized trees produced less tbliage per centimeter of twig length f<strong>or</strong> the period<br />

1989 through 1992 and buds also were smaller in all years except 1989.<br />

Radial growth usually is related to foliage quantity and quality. Fertilized trees had less foliage production but m<strong>or</strong>e<br />

radial growth than control trees, which suggested that resources may have been allocated to the tree bole at the expense of<br />

foliage. Fertilization resulted in greater individual tree growth similar to findings of Cochran (1978). He fertilized thinned<br />

ponderosa pine stands similar and close to our study area; however, his stands were not defoliated by pand<strong>or</strong>a moth at that<br />

time.<br />

A study by Miller and Wagner (1989) of the effects of pand<strong>or</strong>a moth defoliation on ponderosa pine growth in Arizona<br />

found greater radial growth in heavily defoliated trees l year after the last defoliation. This is not the usual growth response<br />

of trees defoliated by some other insects where growth is usually directly related to degree of defoliation and lags 1 year after<br />

defoliation (Wickman 1979, Wickman etal. in press)<br />

The study raised m<strong>or</strong>e questions than answers. Response of insects and foliage production of host trees was contrary<br />

to findings from a recent study of western spruce budw<strong>or</strong>m (Mason et al. 1992, Waring et al. 1992, Wickman et hi. 1992).<br />

But radial growth seems to respond positively after fertilization on nutrient deficient soils irrespective of insect <strong>or</strong> host tree<br />

species, and perhaps from a site productivity perspective that is what really matters.<br />

It may be impossible to account f<strong>or</strong> the smaller larvae from fertilized trees in <strong>this</strong> small study. Perhaps the added<br />

nutrient resources (N) were only partially allocated f<strong>or</strong> growth of trees and some increased defensive chemistry production in<br />

fertilized trees resulted in smaller larvae. Maybe the larval growth was related to feeding efficiency determined by effects of<br />

foliage phenolics <strong>or</strong> other foliage attributes, like needle toughness, on the previous generation of larvae that carried over to<br />

the next generation. This explanation is difficult to pursue because we were able to measure larval weights from only one<br />

generation of the three that affected the host tree through the outbreak period. Perhaps the insect frass with its high levels of<br />

available nitrogen reacted synergistically with residues of N in the soil and <strong>this</strong> resulted in increased radial growth, even<br />

while the trees were being defoliated. Miller and Wagner (1989) suggest that the heavily defoliated pines ability to compensate<br />

f<strong>or</strong> foliage loss involves certain compensat<strong>or</strong>y growth mechanisms, perhaps accelerated nutrient cycling of insect frass.<br />

We did find that pand<strong>or</strong>a moth frass is rich in nutrients, but we had no way of accounting f<strong>or</strong> interactions of frass and our N<br />

treatments.<br />

This study, though small and sh<strong>or</strong>t on chemical analysis, does point out the vagaries associated with interactions<br />

among fertilization, insects, defoliation, and the host tree. Our studies to date of budw<strong>or</strong>m on fir and pand<strong>or</strong>a moth on pine<br />

do not indicate that fertilization enhances plant defenses against herbiv<strong>or</strong>es through the production of secondary resistance<br />

compounds. One consistent result we have encountered is increased radial growth of fertilized trees.<br />

SUMMARY<br />

Responses of ponderosa pine and pand<strong>or</strong>a moth to fertilization with N were studied f<strong>or</strong> 4 years after treatment.<br />

Fertilization had a negative effect on larval weights. The 1990 generation of treated larvae were significantly smaller than<br />

control larvae, but there was no significant difference between pupal weights. Foliage and bud weights of fertilized trees<br />

were significantly lighter than controls. Available nitrogen in both foliage and insect frass was higher in fertilized plots.<br />

Radial growth at dbh of fertilized trees was significantly greater and almost double the growth of controls. The results<br />

indicate that effects of fertilization differ with species of host tree and insect herbiv<strong>or</strong>e, except f<strong>or</strong> increased radial growth<br />

that has been consistently noted in all the recent fertilization studies in eastern Oregon. The highly complex interactions of<br />

increased nutrient cycling from herbiv<strong>or</strong>e feeding and artificial application of N may help explain variable results found in<br />

recent studies.<br />

ACKNOWLEDGMENTS<br />

Technical assistance was provided by Katie Bobowski, Ellen Stenard, and Wendy Sutton. We also thank Dr.<br />

Russel Mitchell f<strong>or</strong> collecting insect frass and Dr. Arthur Tiedemann f<strong>or</strong> arranging f<strong>or</strong> chemical analysis of foliage and frass.<br />

125


Dr. William Mattson arranged f<strong>or</strong> analysis of sugars and phenolics in foliage. Valuable insights which helped our<br />

interpretations were provided by Drs. William Mattson and Pekka Niemelfi during a visit to the study site in 1993. We<br />

appreciate helpful review comments from Dr. Arthur Tiedemann and Dr. Patrick Cochran, Pacific N<strong>or</strong>thwest <strong>Research</strong><br />

<strong>Station</strong>, and Dr. Pekka NiemelS., Finnish F<strong>or</strong>est <strong>Research</strong> Institute.<br />

LITERATURE CITED<br />

CAROLIN, V.M. and KNOPE J.A.C. 1968. The pand<strong>or</strong>a moth. U.S. Dept. Agric., F<strong>or</strong>. Serv. Pest Leafl. 114.<br />

COCHRAN, RH. 1978. Response of a pole-size ponderosa pine stand to nitrogen, phosph<strong>or</strong>us, and sulfur. Res. Note PNW-<br />

319. P<strong>or</strong>tland, OR: U.S. Department of Agriculture, F<strong>or</strong>est Service, Pacific N<strong>or</strong>thwest <strong>Research</strong> <strong>Station</strong>. 8 p.<br />

MASON, R.R., WICKMAN, B.E., BECKWITH, R.C., and PAUL, H.G. 1992. Thinning and nitrogen fertilization in a grand<br />

fir stand infested with western spruce budw<strong>or</strong>m. Part I: Insect response. F<strong>or</strong>. Sci. 38(2):235-251.<br />

MATTSON, W.J., HERMS, D.A., WITTER, J.A., and ALLEN, D.C. 1991. Woody plant grazing systems: N<strong>or</strong>th American<br />

outbreak foliv<strong>or</strong>es and their host plants, p. 53-84. In Baranchikov, Y., Mattson, W.J., Hain, F.R, and Payne, T.L., eds.<br />

F<strong>or</strong>est insect guilds: patterns of interaction with host plants. GTR-NE-153. Radn<strong>or</strong>, PA: U.S. Department of<br />

Agriculture, F<strong>or</strong>est Service. 400 p.<br />

MILLER, K.K. and WAGNER, M.R. 1989. Effect of pand<strong>or</strong>a moth (Lepidoptera:Saturniidae) defoliation on growth of<br />

ponderosa pine in Arizona. J. Econ. Entomol. 82(6): 1682-1686.<br />

PATTERSON, J.E. 1929. The pand<strong>or</strong>a moth, a periodic pest of western pine f<strong>or</strong>ests. U.S. Dept. Agric. Tech. Bull. 137.20 p.<br />

SCHMID, J.M., BENNETT, D., YOUNG, R.W., MATA, S., ANDREWS, M., and MITCHELL, J. 1982. Sampling larval<br />

populations of the pand<strong>or</strong>a moth. F<strong>or</strong>. Res. Note RM-421. U.S. Department of Agriculture, F<strong>or</strong>est Service. 5 p.<br />

VOLLAND, L.A. 11976. Plant communities of the central Oregon pumice zone. R-6 Area Guide 4-2. U.S. Department of<br />

Agriculture, F<strong>or</strong>est Service. 113 p.<br />

WARING, R.H., SAVAGE, T., CROMACK, K., Jr., and ROSE, C. 1992. Thinning and nitrogen fertilization in a grand fir<br />

stand infested with western spruce budw<strong>or</strong>m. Part IV: An ecosystem/pest management perspective. F<strong>or</strong> Sci. 38(2):<br />

275-286.<br />

WICKMAN, B.E. 1979. How to estimate defoliation and predict tree damage. Douglas-fir tussock moth handbook. Agric.<br />

Handb. 550. Washington, DC: U.S. Deptartment of Agriculture. 15 p.<br />

WICKMAN, B.E., MASON, R.R., and PAUL, H.G. 1992. Thinning and nitrogen fertilization in a grand fir stand infested<br />

with western spruce budw<strong>or</strong>m. Part II: Tree growth response. F<strong>or</strong>. Sci. 38(2): 252-264<br />

WICKMAN, B.E., MASON, R.R., and SWETNAM, T.W. (In press). Searching f<strong>or</strong> Long-term patterns of f<strong>or</strong>est insect<br />

outbreaks. In Eather, S., Walters, K., Mills, N., and Watt, A., eds. Proc. Individuals, Populations, and Patterns in<br />

Ecology. L Intercept Press, United Kingdom.<br />

126


PHYTOCHEMICAL PROTECTION AND NATURAL ENEMIES IN THE<br />

REGULATION OF A ROOT AND STEM BORING INSECT<br />

A.V. WHIPPLE and D.R. STRONG<br />

Bodega Marine Lab<strong>or</strong>at<strong>or</strong>y, University of Calif<strong>or</strong>nia, Box 247, Bodega Bay, Calif<strong>or</strong>nia 94923, USA<br />

INTRODUCTION<br />

Much debate has centered on why herbiv<strong>or</strong>es in natural systems do not m<strong>or</strong>e often have a large negative effect on the<br />

plants on which they feed. Except f<strong>or</strong> some aquatic trophic cascades in which plants are algae, (Carpenter et aL 1987,<br />

Carpenter and Kitchell 1988) and a few large mammalian grazers (McNaughton et al. 1988), herbiv<strong>or</strong>es have rarely been<br />

shown to substantially depress plant populations. Two main ideas why <strong>this</strong> is are top-down suppression by predat<strong>or</strong>s and<br />

bottom-up plant protection (Hunter and Price 1992). The natural enemy hypothesis suggests that populations are generally .<br />

controlled by their predat<strong>or</strong>s and/<strong>or</strong> pathogens (Hairston et al. 1960, Oksanen et al. 1981, Hairston and Hairston 1993). Thus,<br />

the plant feeders are kept too sparse to substantially depress plant populations. The endogenous plant protection hypothesis<br />

proposes that, in terrestrial and some aquatic systems, plant biomass is protected by noxious secondary chemicals (Ehrlich<br />

and Raven 1964, Hawkins 1992, Hay and Steinberg 1992). An adjunct to <strong>this</strong> is that higher plant biomass is also largely<br />

unavailable to most herbiv<strong>or</strong>es (Strong 1992, White 1993). This is because most herbiv<strong>or</strong>es cannot digest cellulose <strong>or</strong> lignin<br />

without the aid of microbial symbionts (Martin 1991). Because cellulose and lignin make up a large p<strong>or</strong>tion of the plant,<br />

many essential nutrients f<strong>or</strong> herbiv<strong>or</strong>es are present only in low concentrations, which may limit the growth of herbiv<strong>or</strong>e<br />

populations so that they are not able to have substantial impact on plant populations.<br />

Of course, these hypotheses are not mutually exclusive, and in the system discussed here both plant defense and<br />

herbiv<strong>or</strong>e natural enemies may play a role in determining the level of herbiv<strong>or</strong>e effect on the plant population (Price 1992).<br />

The system of the bush lupine, Lupinus arb<strong>or</strong>eus, and its stem and root b<strong>or</strong>ing ghost moth <strong>or</strong> "swift," Hepialus caI(f<strong>or</strong>nicus,<br />

is ideal f<strong>or</strong> examining the roles of these different fact<strong>or</strong>s because the level of herbiv<strong>or</strong>e attack varies over space with some<br />

areas exhibiting extensive plant m<strong>or</strong>tality and others with little m<strong>or</strong>tality (Strong et al. in prep.).<br />

Directly <strong>or</strong> indirectly, abiotic fact<strong>or</strong>s are probably involved in the differences between the sites in level of attack. The<br />

effect of an abiotic fact<strong>or</strong> on the level of herbiv<strong>or</strong>y may act directly on the herbiv<strong>or</strong>e <strong>or</strong> be mediated through the herbiv<strong>or</strong>e's<br />

host plant, which would be bottom-up regulation, <strong>or</strong> through predat<strong>or</strong>s and pathogens, which would be top-down suppression.<br />

Herbiv<strong>or</strong>y<br />

and Resistance in the Bush Lupine<br />

The bush lupine is a dominant woody shrub of some coastal headland habitats in central Calif<strong>or</strong>nia. At Point Reyes<br />

and on Bodega Head in Marin and Sonoma Counties, a maj<strong>or</strong> source of m<strong>or</strong>tality to the bush lupine appears to be the swift,<br />

Hepialus calif<strong>or</strong>nicus (Opler 1968, Davidson and Barbour 1977, Wagner 1986). The mid to late instars of the swift caterpillar<br />

b<strong>or</strong>e in the roots and stem of the bush lupine. The early instars are thought to feed on litter and rootlets (Wagner 1986).<br />

The caterpillars reach 40-50 mm, and as many as 60 have been found in one bush. While those numbers are exceptional, it is :<br />

difficult to find a large bush with no evidence of swift caterpillar damage (pets. obs.).<br />

Lupines produce large quantities of quinolizidine alkaloids, which are thought to function in nutrient st<strong>or</strong>age and in<br />

defense against herbiv<strong>or</strong>es and pathogens (Wink 1992). As little as 0.11%fresh weight of these alkaloids can be toxic to<br />

insect herbiv<strong>or</strong>es (Wink 1992). Bush lupine leaves contain high levels of these quinolizidine alkaloids in their foliage _<br />

(Bentley and Johnson 1991). The bush lupine probably also has substantial levels of alkaloids in the stems and roots where<br />

the swift caterpillars feed (Wink and Witte 1984). The alkaloid is probably concentrated in the epidermis of the stem and root<br />

where exceptionally high levels of alkaloid have been found in other lupines (Wink 1992).<br />

Mattson, W.J., Niemel/i, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

127<br />

!<br />

ii


Whether these quinolizidine alkaloids defend the plant against attack by a particular herbiv<strong>or</strong>e species could depend<br />

on whether the herbiv<strong>or</strong>e in a specialist <strong>or</strong> a generalist. A specialist could be indifferent <strong>or</strong> prefer high levels of alkaloid, and<br />

a generalist would avoid <strong>or</strong> be harmed by high levels of alkaloid. Wink (1992):reviews the evidence f<strong>or</strong> both of these cases;<br />

however, conclusive evidence f<strong>or</strong> effects of alkaloids in nature has not yet been shown (ttartmann 1992). F<strong>or</strong> swifts, the<br />

generalist-specialist distinction is ambiguous. The insect is polyphagous, having been found to feed on a wide variety of host<br />

plants over its geographical range. In addition, it is even m<strong>or</strong>e indiscriminantly polyphagous and fungiv<strong>or</strong>ous in early instars<br />

when feeding on detritus, bef<strong>or</strong>e entering the root. However, as with many polyphagous species, local populations may be<br />

much m<strong>or</strong>e restricted in their host plant use than <strong>this</strong> implies (Fox and M<strong>or</strong>row 1981, Singer and Parmesan 1993). The swifts<br />

at Bodega Marine Reserve seem to be feeding largely on lupine (though perhaps not entirely), despite the presence of hosts<br />

they have been found on in other locations (De Benedictis et al. 1990), and could show characteristics of specialist rather<br />

than generalist herbiv<strong>or</strong>es. Hence, it is not obvious what effect, if any, quinolizidine alkaloids will have on the interaction<br />

between the swifts and the lupine. The same can be said f<strong>or</strong> another class of chemicals thought to function in defense that<br />

have been found in lupine roots, the isoflavonoids (Lane et al. 1987).<br />

Herbiv<strong>or</strong>y, and perhaps levels of endogenous resistance of bush lupine, vary over the reserve on Bodega Head.<br />

Dramatically m<strong>or</strong>e bushes died, and each contained many m<strong>or</strong>e caterpillars, in some areas than in others in 1992 and 1993<br />

(Strong et al. in prep.). In addition, the areas with lower bush lupine m<strong>or</strong>tality in these 2 years have had a fairly consistent<br />

cover of lupines over the past 40 years. Conversely, areas with greater recent m<strong>or</strong>tality show a hist<strong>or</strong>ical pattern of fluctuating<br />

cover (Strong et al. in press). The plants in areas with different levels of herbiv<strong>or</strong>y may differ in their resistance to swift<br />

herbiv<strong>or</strong>y. Thus, in terms of endogenous plant protection, lower levels of herbiv<strong>or</strong>y by swift caterpillars could derive from<br />

lower nutritional value of the plant tissue to the herbiv<strong>or</strong>e, from higher secondary chemical levels, <strong>or</strong> both since in some<br />

cases they are inex<strong>or</strong>ably intertwined.<br />

The bark and epidermis of the lupine may constitute another <strong>or</strong> different defense because the early instars may be<br />

prevented from feeding on the plant by some chemical <strong>or</strong> physical property of the outer bark. Thus, they would be m<strong>or</strong>e<br />

exposed to natural enemies f<strong>or</strong> a longer period of time.<br />

Natural Enemies of Swift Caterpillars<br />

The early instars of Hepialus calif<strong>or</strong>nicus are very small, with a starting length of approximately 1 ram, and could be<br />

quite vulnerable to predation because they feed externally on litter and roots. The high fecundity of the swift, followed by<br />

modest numbers of large caterpillars in lupine stems, indicates high early instar m<strong>or</strong>tality. Potential predat<strong>or</strong>s in the soil<br />

around the roots include a geophilom<strong>or</strong>ph centipede, which does eat early instars in lab<strong>or</strong>at<strong>or</strong>y trials (Beld unpubl, data).<br />

Wagner (1986) has rep<strong>or</strong>ted cannibalism as a maj<strong>or</strong> source of m<strong>or</strong>tality in lab<strong>or</strong>at<strong>or</strong>y colonies as well.<br />

As the caterpillars grow larger they are found feeding externally on lupine roots, at which time swift caterpillars<br />

killed by the new nematode species Heter<strong>or</strong>habditis sp. have been found. Heter<strong>or</strong>habditis sp. probably kills the early instars<br />

as well, but the tiny c<strong>or</strong>pses are simply harder to locate. Larger caterpillars tunnel inside the lupine root and stem where they<br />

become relatively invulnerable to predation. A fungus, Beuvaria brogniartii, kills the medium to large swifts inside the stem<br />

and root (Williams 1905, Wagner 1986, Strong and Kaya unpubl.).<br />

The nematode has the potential to limit the density of swirl caterpillars in the stems. The prevalence of the nematodes<br />

in the soil at the base of lupine sterns, as measured by a GaUeria mellonella bioassay (Bedding and Ackhurst 1975), is<br />

quite variable between sites (Strong unpubl, data).<br />

Abiotic Fact<strong>or</strong>s<br />

Several abiotic fact<strong>or</strong>s may be of imp<strong>or</strong>tance in <strong>this</strong> system either by depressing swirl populations directly <strong>or</strong> by<br />

affecting lupines <strong>or</strong> the nematode. Wind could affect swift moth populations at the different sites directly; ovipositing female<br />

swifts release eggs as they fly over suitable habitat. The prevailing winds come out of the n<strong>or</strong>th, c<strong>or</strong>related with low swirl<br />

caterpillar density in lupine roots. However, Wagner (1986) found that the moths do not usually fly if it is too windy and one<br />

of their flight times is predawn when winds are usually light. Also, desiccation may be a maj<strong>or</strong> cause of m<strong>or</strong>tality in the early<br />

instar caterpillars.<br />

128


Indirect effects of abiotic fact<strong>or</strong>s on swift populations could be mediated either by their host plant <strong>or</strong> through predat<strong>or</strong>s<br />

and pathogens. Input of sea spray is higher at the upwind sites (Barbour et al. 1973). One consequence of higher salt<br />

spray loads could be higher soil phosph<strong>or</strong>ous levels. A possible mechanism by which phosph<strong>or</strong>us could drastically alter<br />

allocation patterns in the plant has been demonstrated by Dinkelaker et al. (1989) in a congeneric species, Lupinus albus.<br />

Under conditions of low phosph<strong>or</strong>us, <strong>this</strong> plant secretes large quantities of citrate (up to 23% of its dry matter production)<br />

from proteoid roots. The citrate enhances the solubility of phosph<strong>or</strong>us, making it m<strong>or</strong>e available to the plant. Thus, plants<br />

that have m<strong>or</strong>e phosph<strong>or</strong>ous available could have much m<strong>or</strong>e fixed carbon to allocate to growth rather than to acquiring<br />

nutrients since they would not need to secrete as much citrate to get their required phosph<strong>or</strong>ous. If a similar mechanism is<br />

present in L. arb<strong>or</strong>eus, <strong>this</strong> would almost certainly alter plant chemistry between sites with different levels of phosph<strong>or</strong>ous in<br />

ways that could affect individual growth rates and survival of the swift caterpillars by changing the nutritional quality of their<br />

food. Differential g_:owth and survival rates could in turn determine the density of caterpillars in the stems and roots of<br />

plants.<br />

Abiotic fact<strong>or</strong>s, such as soil properties and moisture, could contribute to differences in the nematode populations<br />

between the two sites. Heter<strong>or</strong>habditis sp. are probably vulnerable to desiccation (Kaya 1990), hence in the dry summer,<br />

slight differences in soil moisture between sites may be quite imp<strong>or</strong>tant to their survival.<br />

DISCUSSION<br />

The picture sketched by our preliminary inf<strong>or</strong>mation leads to several hypotheses concerning the differences in the<br />

impact of herbiv<strong>or</strong>e on its host plant among our different sites. Understanding the mechanism by which plants are "protected"<br />

from herbiv<strong>or</strong>y at sites with low herbiv<strong>or</strong>y as compared to sites with high herbiv<strong>or</strong>y will allow us to assess the<br />

relative imp<strong>or</strong>tance of the top-down effects of natural enemies and the bottom-up effects of plant chemistry (including<br />

nutrition). The simplest hypotheses implicate a single abiotic fact<strong>or</strong> <strong>or</strong> an abiotic fact<strong>or</strong> mediated by either the nematode <strong>or</strong><br />

the lupine. These include differences in oviposition due to wind, differences in levels of predat<strong>or</strong>y nematodes due to differences<br />

in soil chemistry, and differences in the secondary chemistry of the lupine due to differences in nutrient levels. While<br />

each of these may play a role, it seems unlikely that any of them are sufficient alone.<br />

An additional hypothesis involves the interaction of plant chemistry and predation in explaining the observed pattern.<br />

The hypothesis is that the different nutrient levels at some sites allow plants to be better defended against herbiv<strong>or</strong>y, <strong>or</strong> to be<br />

m<strong>or</strong>e vig<strong>or</strong>ous and better able to withstand herbiv<strong>or</strong>y (Price 1991), both of which would reduce the impact of the herbiv<strong>or</strong>e<br />

enough that the plants would be able to persist at that site. Then, since the plants persist at that site, rather then undergoing<br />

local extinction, there is a constant supply of the swift f<strong>or</strong> the nematode to prey upon. At sites that experience local extinction<br />

of lupine, the nematode could have no host f<strong>or</strong> a period of time and go extinct at that site. Since it is not a very mobile<br />

predat<strong>or</strong>, it may not be able to recolonize rapidly when the lupine and the swift reappear. Thus, the nematode reinf<strong>or</strong>ces the<br />

pattern caused by the differences in herbiv<strong>or</strong>e "resistance" between the sites by reducing swift populations where the lupines<br />

usually persist, but not where they experience high m<strong>or</strong>tality.<br />

Experimental manipulations in the field of natural enemies and conditions affecting plant chemistry will be needed to<br />

assess the relative roles of top-down control and bottom-up protection in controlling herbiv<strong>or</strong>e populations. The third<br />

alternative, as we have mentioned, is the action of abiotic fact<strong>or</strong>s directly on the herbiv<strong>or</strong>e populations; <strong>this</strong> possibility should<br />

also be examined by looking at egg deposition and the desiccation of early instars at the different sites.<br />

SUMMARY<br />

Mature bush lupine, Lupinus arb<strong>or</strong>eus, die after heavy stem and root b<strong>or</strong>ing by caterpillars of the ghost moth <strong>or</strong><br />

swift, Hepialus calif<strong>or</strong>nicus. M<strong>or</strong>tality rates of these plants can be extremely high in some stands and quite low in other<br />

stands in close proximity. This system is ideal f<strong>or</strong> testing f<strong>or</strong> top-down versus bottom-up control of herbiv<strong>or</strong>e populations by<br />

examining the differences between sites where there is heavy herbiv<strong>or</strong>y and sites where there is not heavy herbiv<strong>or</strong>y. Nutrient<br />

levels that are likely to affect lupine plant allocation patterns probably vary over the area of interest. Population levels of<br />

an entomopathogenic nematode that prey upon the swift moth also vary over the reserve. It is hypothesized that both plant<br />

chemistry variation and the nematode natural enemy play an imp<strong>or</strong>tant role in determining the level of herbiv<strong>or</strong>e attack.<br />

129


ACKNOWLEDGMENTS<br />

We would like to thank John Maron f<strong>or</strong> access to his data on lupine fluctuations, which is in manuscript. F<strong>or</strong><br />

financial supp<strong>or</strong>t we would like to thank Barbara Bentley f<strong>or</strong> a graduate student grant from the Bodega Field Conferences<br />

(AVW), the Sippl fund f<strong>or</strong> travel supp<strong>or</strong>t (AVW), and William Mattson and the IUFRO conference f<strong>or</strong> aid with<br />

meeting fees (AVW).<br />

LITERATURE CITED<br />

BARBOUR, M.G., CRAIG, R.B., DRYSDALE, F.R., and GHILSEN, M.T. 1973. Coastal Ecology: Bodega Head. University<br />

of Calif<strong>or</strong>nia Press, Berkeley.<br />

BEDDING, R.A. and ACKHURST, R.J. 1975. A simple technique f<strong>or</strong> the detection of insect parasitic rhapbditid nematodes<br />

in soil. Nematologica 21: 109-114.<br />

BENTLEY. B.L. and JOHNSON, N.D. 1991. Plants as tbod f<strong>or</strong> herbiv<strong>or</strong>es: the roles of nitrogen fixation and carbon<br />

dioxide enrichment, p. 257-272. In Price, P.W., Lewinsohn, T.M., Fernandes, G.W., and Benson, W.W., eds. Plant-<br />

Animal Interactions. John Wiley & Sons, New Y<strong>or</strong>k.<br />

CARPENTER, SR. and KITCHELL, J.F. 1988. Consumer control of lake productivity. Bioscience 38: 764-769.<br />

CARPENTER, S.R., KITCHELL, J.F., HODGSON, J.R., COCHRAN, P.A., ELSER, J.J., ELSER, M.M., LODGE, D.M.,<br />

KRETCHER, D., HE, X., and VON ENDE, C.N. 1987. Regulation of Rakeprimary productivity by food web<br />

structure. Ecology 68: 1863-1876.<br />

DAVIDSON, ED. and BARBOUR, M.G. 1977. Germination, establishment, and demography of coastal bush lupine<br />

(Lupinus arb<strong>or</strong>eus') at Bodega Head, Calif<strong>or</strong>nia. Ecology 58: 592-600.<br />

DE BENEDICTIS, J.A., WAGNER, D.L., and WHITFIELD, J.B. 1990. Larval hosts of the microlepidoptera of the San<br />

Bruno Mountains, Calitbrnia. Atala 16: 14-35.<br />

DINKELAKER, B., ROMHELD, V., and MARSCHNER, H. 1989. Citric acid excretion and precipitation of calcium citrate<br />

in the rhizosphere of white lupine (Lupinus albus L.). Plant. Cell. Environ. 12: 285-292.<br />

EHRLICH, PR. and RAVEN, P.H. 1964. Butterflies and plants: a study in coevolution. Evolution 18: 586-608.<br />

FOX, LR. and MORROW, P.A. 1981. Specialization: species property <strong>or</strong> local phenomenon? Science 221: 887-893.<br />

HAIRSTON, N.G., JR and HAIRSTON, N.G., SR. 1993. Cause-effect relationships in energy flow, trophic structure, and<br />

interspecific interactions. Am. Nat. 142: 379-411.<br />

HAIRSTON, N.G., SMITtt, EE., and SLOBODKIN, L.B. 1960. Community structure, population control, and competition.<br />

Am. Nat. 94:421-424.<br />

HARTMANN, T. 1992. Alkaloids, p. 79-116. In Rosenthal, G.A. and Berenbaum, M.R., eds. Herbiv<strong>or</strong>es: Their Interactions<br />

with Secondary Plant Metabolites. Second edition, vol. I. Academic Press, San Diego.<br />

HAWKINS, B.A. 1992. Parasitoid-host food webs and don<strong>or</strong> control. Oikos 65: 159-162.<br />

HAY, M.E. and STEINBERG, ED. 1992. The chemical ecology of plant-herbiv<strong>or</strong>e interactions in marine versus terrestrial<br />

communities, p. 372-408. In Rosenthal, G.A. and Berenbaum, M.R., eds. Herbiv<strong>or</strong>es: Their Interactions with<br />

Secondary Plant Metabolites. Second edition, vol. I. Academic Press, San Diego.<br />

130


HUNTER, M.D. and PRICE, P.W. 1992. Playing chutes and ladders: heterogeneity and the relative roles of bottom-up and<br />

top-down f<strong>or</strong>ces in natural communities. Ecology 73: 724-732.<br />

KAYA, H.K. t990. Soil ecology, p. 93-11 I. In Gaugler, R. and Kaya, H.K., eds. Entomopathogenic Nematodes in Biological<br />

Control. CRC Press, Boca Raton.<br />

LANE, G.A., SUTHERLAND, O.R.W., and SKIPP, R.A. 1987. Isoflavonoids as insect feeding deterrents and antifungal<br />

components from root of Lupinus augustijblius. J. Chem. Ecol. 13" 771-783.<br />

MARTIN, M.M. 1991. The evolution of cellulose digestion in insects. Phil. Trans. R. Soc. Lond. B. 333: 281-288.<br />

MCNAUGHTON, S.J., RUESS, R.W., and SEAGLE, S.W., 1988. Large mammals and process dynamics in African ecosystems.<br />

Bio Science 38 794-800.<br />

OKSANEN, L., FRETWELL, S.D., ARRUDA, J., and NIEMELA, P. 1981. Exploitation ecosystems in gradients of primary<br />

productivity. Am. Nat. 118" 240-261.<br />

OPLER, RA. 1968. Unusual numbers of Hepialus sequoiolus Behrens in Sonoma County. Pan. Pac. Entomol. 44: 83.<br />

PRICE, RW. 1991. The plant vig<strong>or</strong> hypothesis and herbiv<strong>or</strong>e attack. Oikos 62:244-251.<br />

PRICE P.W. 1992. Plant resources as the mechanistic basis f<strong>or</strong> insect herbiv<strong>or</strong>e population dynamics, p. 139-173. In Hunter,<br />

M.D., Ohgushi, T., and Price, RW., eds. Effects of Resource Distribution on Animal-Plant Interactions. Academic<br />

Press, San Deigo.<br />

SINGER, M.C. and PARMESAN, C. 1993. Sources of variations in patterns of plant-insect association. Nature 361:251-<br />

253.<br />

STRONG, D.R. 1992. Are trophic cascades all wet? Differentiation and don<strong>or</strong>-control in speciose ecosystems. Ecology 73:<br />

747-754.<br />

STRONG, D.R., MARON, J.L., and CONNORS RG. 1996. Phmt competition, endogenous defenses, and natural enemies in<br />

terrestrial food webs: an example from the bush lupine community. In Polls, G. and Winemiller, K., eds. Food Webs:<br />

integration of patterns and dynamics. Chapman and Hall, NY. 472 p.<br />

STRONG, D.R., MARON, J.L., HARRISON, S., CONNORS, RG., and JEFFRIES, R.L. 1995. High m<strong>or</strong>tality, fluctuation<br />

in numbers, and heavy subterranean insect herbiv<strong>or</strong>y in bush lupine, Lupinus arb<strong>or</strong>eus, ms to be submitted to<br />

Oecologia.<br />

WAGNER, D.L. 1986. The biosystematics of Hepialus F.s. lato with special emphasis on the calif<strong>or</strong>nicus-hectoides species<br />

group. PhD. Thesis in Entomology, University of Calif<strong>or</strong>nia, Berkeley.<br />

WILLIAMS, F.X. 1905. Notes on the life hist<strong>or</strong>y of Hepialus sequoiolus Behrens. Entomol. News. 16: 283-287.<br />

WHITE, T.C.R. 1993. The Inadequate Environment: Nitrogen and the Abundance of Animals. Springer-Verlag, New Y<strong>or</strong>k.<br />

WINK, M. 1992. The role of quinolizidine alkaloids in plant-insect interactions, p. 131-166. In Bernays, E., ed. Plant-Insect<br />

Interactions. vol. IV. CRC Press, Boca Raton, FI.<br />

WINK, M. and WITTE, L. 1984. Turnover and transp<strong>or</strong>t of quinolizidine alkaloids. Diurnal fluctuations of lupanine in the<br />

phloem sap, leaves, and fruits ofLupinus albus L. Planta 161: 519-524.<br />

131


RESISTANCE OF HYBRID AND PARENTAL WILLOWS TO HERBIVORES:<br />

HYPOTHESES AND VARIABLE HERBIVORE RESPONSES OVER 3 YEARS<br />

ROBERT S. FRITZ l, BERNADETTE M. ROCHE I, and COLIN M. ORIANS 2<br />

_Department of Biology, Vassar College, Poughkeepsie, New Y<strong>or</strong>k 12601, USA<br />

2Department of Biology, Williams College, Williamstown, Massachusetts 01267, USA<br />

INTRODUCTION<br />

Analysis of herbiv<strong>or</strong>y on hybrid plants in natural populations has recently gained attention (Whitham 1989, Keim et<br />

al. 1989, Boecklen and Spellenberg 1990, Paige et al. 1991, Whitham et al. 1991, Aguilar and Boecklen 1992, Fritz et al.<br />

1994, Whitham et al. 1994), and it may be imp<strong>or</strong>tant f<strong>or</strong> understanding plant-herbiv<strong>or</strong>e interactions (Whitham 1989). To<br />

assess the significance of interspecific hybridization of plants f<strong>or</strong> plant-herbiv<strong>or</strong>e interactions, it will be imp<strong>or</strong>tant to know<br />

how interspecific hybridization of plants influences resistance of hybrid plants in relation to resistance of pure parents. It has<br />

been hypothesized that hybrid plants may act as an "ecological sink" f<strong>or</strong> herbiv<strong>or</strong>ous insects (Whitham 1989). In mixed<br />

stands of parental and hybrid plants, <strong>this</strong> hypothesis predicts that a large fraction of the herbiv<strong>or</strong>e population would reside on<br />

hybrid individuals (Drake 1981, Whitham 1989, Floate et al. 1993). Whitham (1989) referred to highly susceptible hybrid<br />

poplars as being an ecological sink f<strong>or</strong> a galling aphid, but in terms of metapopulation dynamics, hybrid plants could function<br />

as a "source" rather than a sink (e.g., Ericson et al. 1993). Hybrid plants have also been hypothesized to act as "evolutionary<br />

sinks" to pests (Whitham 1989). The "evolutionary sink" idea predicts that if herbiv<strong>or</strong>e fitness was higher on hybrids<br />

compared to their fitness on parental plants, there would be selection f<strong>or</strong> specialization on the hybrids, perhaps preventing<br />

selection tbr herbiv<strong>or</strong>e virulence on the parental species. If so, then hybrid plant-herbiv<strong>or</strong>e interactions may be critical to<br />

understanding why plants seem to maintain pest resistance in the face of a greater evolutionary potential f<strong>or</strong> virulence in<br />

pests. Finally, Floate and Whitham (1993) suggest that hybrid plants may facilitate host shifts of herbiv<strong>or</strong>es onto new plants,<br />

because they could act as a selective "bridge" fav<strong>or</strong>ing adaptation of an herbiv<strong>or</strong>e to previously non-host species. Tests of<br />

these hypotheses will require knowing the general patterns of herbiv<strong>or</strong>e responses to hybrid versus pure parental species.<br />

Hypotheses<br />

Fritz et al. (1994) proposed several alternative hypotheses of phytophage response to hybrid plants. Those hypotheses<br />

assume that pure parent species are being compared to F1 hybrids. The reason to focus on F1 hybrids is that <strong>this</strong><br />

comparison could suggest: (1) how resistance mechanisms are inherited from parental species, and (2) how these traits may<br />

be expressed upon backcrossing. The Additive hypothesis (Fig. 1A) predicts that hybrids do not differ from the mean of the<br />

resistances of the two parents (i.e., the midparent value). Thus, F1 hybrids are intermediate between the resistances of the<br />

parental species, which suggests that hybrid resistances are due to the additive inheritance of resistance traits from both<br />

parents.<br />

The second hypothesis is the Dominance hypothesis (Fig. 1B). If hybrid resistance differs significantly from the<br />

mean resistance of both parents but does not differ significantly from that of one of the parents, the Dominance hypothesis<br />

would be supp<strong>or</strong>ted. Hybrid resistance could resemble that of either the m<strong>or</strong>e resistant <strong>or</strong> the m<strong>or</strong>e susceptible parent.<br />

(Dominance in <strong>this</strong> context refers to phenotypic similarity between a parent and hybrids, not necessarily genetic dominance,<br />

though it may imply dominant inheritance of resistance traits.) Herbiv<strong>or</strong>e densities could be intermediate between the<br />

Additive and Dominance patterns, which would supp<strong>or</strong>t an hypothesis of partial dominance.<br />

The third hypothesis is the Hybrid Susceptibility hypothesis (the Hybrids-as-Sinks hypothesis of Whitham 1989; see<br />

also Keim et al. 1989, Boecklen and Spellenberg 1990). This hypothesis predicts higher herbiv<strong>or</strong>e densities and/<strong>or</strong> higher<br />

herbiv<strong>or</strong>e pertbrmance on hybrids compared to parental taxa (Fig. 1C). The Hybrid Susceptibility hypothesis can be distinguished<br />

from the Dominance hypothesis in that the susceptibility of the hybrid must exceed that of the most susceptible<br />

parent.<br />

Mattson, W.J., Niemelfi, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

132


A° ADDITIVE C. HYBRID SUSCEPTIBILITY<br />

50 50<br />

)- 40 >- 40<br />

03 03<br />

Z Z<br />

w 30 w<br />

Q 0 30<br />

uJ LU<br />

O 20 O 20<br />

n- nw<br />

10 w<br />

I _: 10<br />

SPECIES 1 HYBRID SPECIES 2 SPECIES 1 HYBRID SPECIES 2<br />

B. DOMINANCE D. HYBRID RESISTANCE<br />

50 50<br />

>- 40 >'- 40<br />

F- __.<br />

03<br />

z z w 30<br />

w 30 o<br />

Q<br />

w LLI<br />

n-" n-<br />

O 2O O 20<br />

n- n-<br />

w 10<br />

w<br />

:z:<br />

10<br />

I<br />

0 "- 0<br />

SPECIES 1 HYBRID SPECIES 2 SPECIES 1 HYBRID SPECIES 2<br />

Figure I. Hypothesized patterns that would result from the Additive hypothesis (A), the Dominance hypothesis (B), the Hybrid<br />

Susceptibility hypothesis (C), and the Hybrid Resistance hypothesis (D). Imaginary densities (Susceptibility) are plot-<br />

ted on the Y-axis f<strong>or</strong> each parent species and their FI interspecific hybrid.<br />

The final hypothesis, the Hybrid Resistance hypothesis, predicts that hybrid plants will be m<strong>or</strong>e resistant than either<br />

parent, resulting in lower herbiv<strong>or</strong>e densities on hybrid plants (Fig. ID). This hypothesis requires that hybrids be m<strong>or</strong>e<br />

resistant than the most resistant parent. A study by Boecklen and Spellenberg (1990) suggests supp<strong>or</strong>t f<strong>or</strong> <strong>this</strong> hypothesis f<strong>or</strong><br />

herbiv<strong>or</strong>e communities in two oak hybrid zones. Hybrids supp<strong>or</strong>ted lower densities and diversities of herbiv<strong>or</strong>es than the<br />

mean of the parental species, but they did not distinguish between the fit of their data to the Hybrid Resistance hypothesis <strong>or</strong><br />

the Dominance hypothesis. We present and discuss these hypotheses in genetic terms since supp<strong>or</strong>t of any of the hypotheses<br />

may suggest probable underlying mechanisms of inheritance of resistance.<br />

It is not clear a pri<strong>or</strong>i why one of these hypotheses should be m<strong>or</strong>e likely than any others. If the Hybrid Susceptibility<br />

hypothesis is prevalent and herbiv<strong>or</strong>y reduces plant fitness, then herbiv<strong>or</strong>es could be imp<strong>or</strong>tant in maintaining distinct plant<br />

species by limiting the fitness of hybrids. Boecklen and Spellenberg (1990) found evidence f<strong>or</strong> the Hybrid Resistance and f<strong>or</strong><br />

133<br />

m


our Additive hypothesis. Fritz et al. (1994) found supp<strong>or</strong>t f<strong>or</strong> the Additive, Dominance, and Hybrid Susceptibility hypotheses<br />

among herbiv<strong>or</strong>e and a pathogen species on hybrid and parental willows. The purposes of <strong>this</strong> paper are to investigate the<br />

responses of I 1 herbiv<strong>or</strong>e species from different guilds to parental and hybrid willows over three consecutive years, and to<br />

determine if herbiv<strong>or</strong>e responses to hybrid plants vary over time. This paper will show that different herbiv<strong>or</strong>es, even those<br />

in the same guild, vary dramatically in their response to hybrid versus parental taxa, but that most species, when considered<br />

over time, fit the Additive Hypothesis.<br />

METHODS AND MATERIALS<br />

The System<br />

This system has several advantages f<strong>or</strong> studying herbiv<strong>or</strong>e-hybrid plant interactions. The two willow species that<br />

hybridize often co-occur in the same habitats with their hybrids, eliminating, to a certain extent, confounding environmental<br />

variation across naturally occurring hybrid zones with herbiv<strong>or</strong>e resistance variation among hybrids and parents. Willows<br />

have a number of herbiv<strong>or</strong>e species in different guilds, including gall-f<strong>or</strong>mers, leaf miners, leaf tiers, and chewers, that are<br />

welt characterized and that have been studied extensively on one of the willow species in <strong>this</strong> system.<br />

Salix sericea Marshall and Salix. eriocephala Michx. co-occur in swamps and along streams in central New Y<strong>or</strong>k. S.<br />

sericea is a 0.5-4-m-high shrub that has lanceolate leaves with densely sericeous hairs on the lower leaf surface. Stipules are<br />

small, lanceolate, and usually absent from older leaf nodes (i.e., stipules are deciduous). This species occurs predominantly<br />

in swamps from Canada south through the n<strong>or</strong>theastern U.S. and along the Appalachian range to Ge<strong>or</strong>gia (Argus 1986). S.<br />

eriocephala is a shrub that may reach 6 m in height, and its leaves are lanceolate to narrowly oblong and glabrous beneath.<br />

Stipules are large, persistent, half ovate, and half c<strong>or</strong>date at the base. It frequently occurs along streams, and its range<br />

extends from Canada south as far as Virginia and west to Missouri (Argus 1986). The ranges of the two species are broadly<br />

sympatric, and at the study site both species co-occur and intermingle. Species can be easily distinguished in the field based<br />

on leaf, stipule, and bud characteristics (Argus 1986).<br />

These willow species hybridize to f<strong>or</strong>m plants that are usually distinctive from each parent. Large, persistent, halfovate<br />

stipules (a S. eriocephala trait) and a sericeous (hairy) lower leaf surface (a S. sericea trait) serve as m<strong>or</strong>phological<br />

markers to identify S. sericea x S. eriocephala hybrids in the field (Fritz et al. 1994). The natural hybrids at the study site do<br />

not appear to be extra<strong>or</strong>dinary; S. eriocephala and S. sericea have been rep<strong>or</strong>ted to hybridize throughout their range (Argus<br />

1974, t986; Mosseler and Papadopol 1989). Flowering phenology may explain the occurrence of hybridization between S.<br />

sericea and S. eriocephala (Argus 1974, Mosseler and Papadopol 1989), which differ by only a few days in the onset of<br />

flowering (foliage phenology is also similar) (Fritz, pets. obs.).<br />

Many herbiv<strong>or</strong>es attack the parental and hybrid willows. Each herbiv<strong>or</strong>e species attacks both parental willow species<br />

and hybrids. Gall-f<strong>or</strong>ming sawflies (Hymenoptera: Tenthredinidae) include Phyllocolpa nigrita, Phyllocolpa sp. nov.,<br />

Phyllocolpa terminalis, and Pontania sp. (leaf galls). Gall-f<strong>or</strong>ming flies (Diptera: Cecidomyiidae) include the stem gallers<br />

Rabdophaga rigidae, the beaked willow gall; R. salicisbrassicoides, the willow cabbage gall; and a leaf galler, Iteomyia<br />

salic(folia. The other common leaf galler species is the gall mite Aculops tetanothrix (Acarina: Eriophyidae). There are two<br />

species of leaf miners, Phyllon<strong>or</strong>ycter salicifoliella and Phyllocnistis sp. (Lepidoptera: Gracillariidae), that commonly f<strong>or</strong>m<br />

mines on willow leaves.<br />

There are two frequently seen leaf folder species (Lepidoptera) that have distinctive leaf folds, but have not yet been<br />

successfully reared <strong>or</strong> identified. One species (LF) folds over and sews the leaf margin to the lamina with silk. The other<br />

species (V) f<strong>or</strong>ms a tube of the leaf tip by sewing the edges of the leaf blade together.<br />

Methods<br />

These studies were perf<strong>or</strong>med from 1991 to 1993 at the Sosnowski site 3 km west of Milf<strong>or</strong>d, NY, along County<br />

Route 44. All plants are individually marked at the study site. Censuses of herbiv<strong>or</strong>es species on 14 S. eriocephala, 29 S.<br />

sericea, and 16 hybrids were conducted from late July to early August 1991. In 1992, we censused 20 S. eriocephala, 36 S.<br />

sericea, and 39 hybrids, and in 1993 we censused 40 S. eriocephala, 109 S. sericea, and 38 hybrids. Mostly hybrid plants<br />

known to be Fl-types (Fritz et al. 1994) were included in the hybrid categ<strong>or</strong>y. Plants known to be backcrosses from RAPD<br />

were not included in <strong>this</strong> analysis. Some hybrid plants added to the census group after 1991 have not been examined f<strong>or</strong> their<br />

RAPD genotype, and it is theref<strong>or</strong>e possible that they are something other than Fl-types (Fritz et al. 1994).<br />

134


Galls, leaf miners, <strong>or</strong> leaf folds were counted on 50 shoots per plant f<strong>or</strong> hybrids and S. eriocephala and on 300 shoots<br />

f<strong>or</strong> S. sericea in 1991 only. Data are expressed as number of herbiv<strong>or</strong>es per 300 shoots. Because leaf folds of Phyllocolpa<br />

r_igritaand Phyllocolpa sp. nov. were not distinguished on S. sericea in 1991, we combined them f<strong>or</strong> analyses in all years.<br />

All other species densities are considered separately. Species densities were considered statistically independent among<br />

plants within taxa f<strong>or</strong> all years (Fritz et al. 1994), and species-wide significance tests rather than table-wide significance tests<br />

were applied. ANOVAs were perf<strong>or</strong>med f<strong>or</strong> each year separately. We first perf<strong>or</strong>med a multivariate ANOVA, using loge<br />

transf<strong>or</strong>med densities of alt herbiv<strong>or</strong>e species, and then we perf<strong>or</strong>med univariate ANOVAs f<strong>or</strong> each species separately.<br />

Orthoganol contrasts were perf<strong>or</strong>med to test f<strong>or</strong> the fit of each herbiv<strong>or</strong>e species to the hypotheses. We perf<strong>or</strong>med<br />

only two <strong>or</strong>thoganol contrasts to test the significance of herbiv<strong>or</strong>e densities on willow taxa (species) to avoid overparameterization<br />

of the analysis, since only three taxonomic groups were present (SAS Institute 1985). To test the Additive hypothesis,<br />

we tested the significance of the hybrid-midparent contrast (Contrast A) using sequential Bonferroni analysis with specieswide<br />

significance at P < 0.05/2 = 0.025 (Rice 1989). If <strong>this</strong> test was significant, we rejected the Additive hypothesis and then<br />

tested the contrast between herbiv<strong>or</strong>e density on hybrids and density on the numerically closest parent (Contrast B) (P < 0.05/<br />

1 = 0.05). This contrast tested if there was a significant departure from the Dominance hypothesis. If the Dominance<br />

hypothesis was rejected, we inspected the means to determine if the Hybrid Susceptibility hypothesis (density on hybrids<br />

exceeded that of the highest parent) <strong>or</strong> the Hybrid Resistance hypothesis (density on hybrids was lower than the lowest<br />

parent) was supp<strong>or</strong>ted, <strong>or</strong> if the density of hybrids was intermediate between that of one parent and the midparent value. This<br />

last possibility would supp<strong>or</strong>t a partial Dominance hypothesis.<br />

RESULTS<br />

Multivariate ANOVAs showed highly significant contrasts between herbiv<strong>or</strong>e densities on hybrids and the mean<br />

herbiv<strong>or</strong>e densities of parent species f<strong>or</strong> all three years (Table 1). This is a rejection of the Additive hypothesis overall.<br />

Further multivariate contrasts could not be made since the test of the Dominance hypothesis requires comparison to the<br />

numerically nearest parent, which would not be the same f<strong>or</strong> all herbiv<strong>or</strong>es. In 1991, five herbiv<strong>or</strong>e species had significant<br />

hybrid vs. parents contrasts (Contrast A) at P < 0.025, and two herbiv<strong>or</strong>e species had marginally significant hybrid vs. parents<br />

contrasts at P < 0.05 (Table 1). In 1992, four species had significant hybrid vs. parents contrasts (Table 1). Three species that<br />

had significant contrasts in 1991 did not have significant contrasts in 1992. In 1993, five species had significant hybrid vs.<br />

parent contrasts (Table 1). Overall, then, there appear to be differences among the years in the species of herbiv<strong>or</strong>es that had<br />

significant departures from the Additive hypothesis.<br />

To illustrate the year-to-year differences in m<strong>or</strong>e detail and to draw conclusions about which hypotheses mentioned<br />

above are supp<strong>or</strong>ted by each species, we present the separate results f<strong>or</strong> each year f<strong>or</strong> each species, which included all<br />

censused plants.<br />

Both contrasts f<strong>or</strong> Phyllonoo, cter salicifoliella were significant in 1991 and 1993 (Table 1), and the means supp<strong>or</strong>t<br />

f<strong>or</strong> the Hybrid Susceptibility hypothesis (Table 2). In I992, neither contrast was significant, which supp<strong>or</strong>ted the Additive<br />

hypothesis. The other leaf miner Phyllocnistis sp. also showed differences between years. In 1991 and 1993, the first<br />

contrast was not significant, supp<strong>or</strong>ting the Additive hypothesis (Tables 1 & 2). Fritz et al. (1994) concluded that <strong>this</strong> species<br />

supp<strong>or</strong>ted the Hybrid Susceptibility hypothesis in 1991, because of the marginally significant contrast. In 1992 the Dominance<br />

hypothesis was supp<strong>or</strong>ted with densities on S. eriocephala and hybrids being equal and lower than on S. sericea (Table<br />

2).<br />

Among species of the leaf galling guild, Phyllocolpa termMatis showed consistent results in each year (Tables I &<br />

2). In all cases the Hybrid Susceptibility hypothesis was supp<strong>or</strong>ted. Phyllocolpa spp. supp<strong>or</strong>ted the Additive hypothesis in<br />

1991 and 1993 (Tables 1 & 2). In each year, including 1992, densities on hybrids were intermediate between the two parental<br />

species. In 1992, the first contrast was significant and the second contrast was not significant, supp<strong>or</strong>ting the Dominance<br />

hypothesis (Tables t & 2). Pontania sp. supp<strong>or</strong>ted the Additive hypothesis in 1991 and 1992, but the Dominance hypothesis<br />

was supp<strong>or</strong>ted in 1993 (Tables 1 & 2). Iteomyia salicifolia supp<strong>or</strong>ted the Hybrid Susceptibility hypothesis in 1991 and 1993<br />

(Tables 1 & 2). In each year, including 1992, densities on hybrids were greater than on either parental species. In 1992, the<br />

first contrast was not significant at P < 0.025 (although it was marginally significant at P < 0.05), thereby supp<strong>or</strong>ting the<br />

Additive hypothesis, even though the mean density on hybrids was m<strong>or</strong>e than twice as high as on S. eriocephala (Table 2).<br />

The last leaf galling species, A. tetanothrix, supp<strong>or</strong>ted the Dominance hypothesis in 1991 and 1993 (Tables 1 & 2). In both of<br />

135


Table 1. --Summary of multivariate and univariate ANOVAs of herbiv<strong>or</strong>e densities on S. sericea, S. eriocephala, and hybrid<br />

plants f<strong>or</strong> 1991, 1992, and 1993. F-values and degrees of freedom are shown f<strong>or</strong> the univariate tests. Contrast A was<br />

considered significant if its P value was less than the Bonferroni criterion of P < 0.025. F-values f<strong>or</strong> the multivariate<br />

ANOVAs are based on Wilks' lambda (other tests gave similar results). Herbiv<strong>or</strong>e species are grouped by feeding<br />

guild.<br />

YEAR 1991 1992 1993<br />

CONTRAST A B A B A B<br />

F(1,61) F(1,61) F(1,92) F(1,92) F(1,184) F(1,184)<br />

Multivariate 11.22? -- 11.28t -- 7.04? --<br />

Leaf Miners<br />

Phyllon<strong>or</strong>ycter<br />

salicifoliella 10.31"** 5.94* 2.06 0.68 7.27** 5.82***<br />

Phyllocnistis sp. 4.24* 4.67* 13.02? 0.22 0.45 5.35*<br />

Leaf Gallers<br />

Phyllocolpa terminalis 18.35t 10.13"** 16.09? 10.26'** 36.91t 27.93?<br />

Phylocolpa spp. 2.18 0.31 6.13"* 0.03 1.66 11.75?<br />

Pontania sp. 0.87 8.60*** 4.28* 0.16 6.86** 1.82<br />

lteomyia salicifolia 13.04? 13.50t 4.32* 0.69 16.95? 11.93?<br />

Aculops tetanothrix 13.97t 2.27 30.23t 7.62*** 11.88t 2.34<br />

Leaf Folders<br />

Leaf folder-LF 4.49* 1.74 0.73 0.40 0.92 0.52<br />

Leaf folder -V 16.99? 9.30*** 0.44 2.36 0.40 0.47<br />

Stem Gallers<br />

Rabdophaga rigidae 1.99 0.21 0.50 0.47 2.13 0.22<br />

salicisbrassicoides 3.24 0.06 4.77* 0.11 0.13 0.44<br />

Total Herbiv<strong>or</strong>es 0.42 9.47*** 1.30 0.32 11.987 4.62*<br />

* - P < .05, ** P < .025, *** P < .01, # - P < .001<br />

these years, densities on S. sericea and hybrids were equal and lower than on S. eriocephala. In 1992, however, the Hybrid<br />

Resistance hypothesis was supp<strong>or</strong>ted. Mite gall density on hybrid plants was significantly lower than on S. sericea. It should<br />

be noted, however, that in all three years the numbers of mite galls were lower on the hybrid plants than on the nearest parent.<br />

F<strong>or</strong> the leaf folding guild, LF supp<strong>or</strong>ted the Additive hypothesis in all years (Tables 1 & 2), but the first contrast was<br />

marginally significant in 1991 (Table 1). Species V supp<strong>or</strong>ted the Hybrid Susceptibility hypothesis in 1991, but the Additive<br />

hypothesis was supp<strong>or</strong>ted in 1992 and 1993 (Tables 1 & 2).<br />

In the stem galling guild, both R. rigidae and R. salicisbrassicoides supp<strong>or</strong>ted the Additive hypothesis consistently in<br />

each year (Tables 1 & 2), but densities of R. salicisbrassicoides in 1992 were marginally significant f<strong>or</strong> the first contrast,<br />

nearly supp<strong>or</strong>ting the Dominance hypothesis.<br />

Seven of the 11 herbiv<strong>or</strong>e species showed significant changes in the hypotheses that were supp<strong>or</strong>ted between years.<br />

In five of the seven cases, densities of the herbiv<strong>or</strong>es had declined between the years of the change; in two cases densities had<br />

increased <strong>or</strong> there was little change. Lower herbiv<strong>or</strong>e densities, theref<strong>or</strong>e, might have contributed to the changes in which<br />

hypotheses were supp<strong>or</strong>ted.<br />

136


Table 2.--Summary of the hypotheses supp<strong>or</strong>ted by each species based on the results in Table 1 and inspection of mean<br />

herbiv<strong>or</strong>e densities among the willow species in each year (Fritz et al. in prep.). When the Dominance hypothesis<br />

was supp<strong>or</strong>ted the two equal taxa are indicated. In all cases they were m<strong>or</strong>e resistant than the other taxon. Herbiv<strong>or</strong>e<br />

species are grouped by feeding guild.<br />

Leaf Miners<br />

1991 1992 1993<br />

Phyllon<strong>or</strong>ycter salicifoliella Susceptibility Additive Susceptibility<br />

Phyllocnistis sp. Additive Dominance (E=ES) Additive<br />

Leaf Gallers<br />

Phyllocolpa terminalis Susceptibility Susceptibility Susceptibility<br />

Phyllocolpa spp. Additive Dominance (ES=S) Additive<br />

Pontania sp. Additive Additive Dominance (E=ES)<br />

lteomyia salicifolia Susceptibility Additive Susceptibility<br />

Aculops tetanothrix Dominance (ES=S) Resistance Dominance (ES=S)<br />

Leaf Folders<br />

Leaf folder-LF Additive ,_dditive Additive<br />

Leaf folder -V Susceptibility Additive Additive<br />

Stem Gallers<br />

Rabdophaga rigidae Additive Additive Additive<br />

R. salicisbrassicoides Additive Additive Additive<br />

Total Herbiv<strong>or</strong>es Additive Additive Susceptibility<br />

To visualize the combined response of herbiv<strong>or</strong>es to hybrids vs. parents, we summed densities of herbiv<strong>or</strong>es on each<br />

plant and reanalyzed the responses. When mites were included in the analysis, there was no significant hybrid vs. parent<br />

contrast in 1991 and 1992 (Tables 1 & 2, Fig. 2), but the Hybrid Susceptibility hypothesis was supp<strong>or</strong>ted in 1993.<br />

DISCUSSION<br />

The main conclusions of <strong>this</strong> study are that: (1) there is evidence that different herbiv<strong>or</strong>es supp<strong>or</strong>t different hypotheses<br />

regarding the effects of interspecific hybridization on plant resistance; and (2) there is year-to-year variation in which<br />

hypothesis is supp<strong>or</strong>ted by a given herbiv<strong>or</strong>e species. These results suggest the possibility that different resistance mechanisms<br />

are inherited differently in hybrid plants, <strong>or</strong> that different herbiv<strong>or</strong>es have different responses to the same resistance<br />

mechanism in the hybrid compared to the parent. The results also suggest the possibility that environmental variation affects<br />

the hybrid and parental willows differently between years, which alters susceptibility to herbiv<strong>or</strong>es.<br />

The hypothesized effects of interspecific hybridization on plant resistance have as an assumption that hybrids are<br />

Fl's. The plants that were censused were predominantly Fl's based on RAPD analysis (Fritz et al. 1994), but six plants were<br />

most closely identified as backcrosses to S. sericea. Known backcross plants were removed from <strong>this</strong> analysis, but it is<br />

possible that some backcrosses are included among the hybrid plants added to the censuses in 1992 and 1993. We have not<br />

yet analyzed these new hybrid plants using our RAPD technique to know their exact genetic status. Fritz et al. (1994) have<br />

shown that m<strong>or</strong>phological characters are not always reliable f<strong>or</strong> distinguishing S. eriocephala from hybrids, and F 1-type<br />

hybrids are indistinguishable from backcross plants using m<strong>or</strong>phological traits (Fritz, unpublished data). With these comments<br />

in mind, we will proceed to interpret the results of <strong>this</strong> study in relation to the hypotheses (Fig. 1).<br />

137


450<br />

TOTAL HERBIVORES<br />

400 [] 1991-]<br />

[] 1992r<br />

350 ID993/<br />

.......<br />

300<br />

250<br />

03<br />

F,, O.<br />

I<br />

IIEI 150<br />

100<br />

50 _,_,_<br />

1/...-/,,, J<br />

O. _.',. .....<br />

ERIOCEPHALA HYBRID SERICEA<br />

SPECIES<br />

Figure 2.--Density (+ 1 SE) of total herbiv<strong>or</strong>e numbers among S. eriocephala, S. sericea and their interspecific hybrids and<br />

among years. Table 2 indicates which hypotheses were supp<strong>or</strong>ted by total herbiv<strong>or</strong>e numbers.<br />

The Additive hypothesis received the greatest supp<strong>or</strong>t in each year (7 of 11 species, Table 2). Herbiv<strong>or</strong>e densities oa<br />

hybrids are often intermediate between the densities on the two parental species. This finding is consistent with an Additive<br />

model of inheritance of resistance traits. Other studies also provide supp<strong>or</strong>t f<strong>or</strong> the Additive hypothesis. Soetens et al.<br />

(1991) found f<strong>or</strong> a species of Pontania on European willows and their hybrids that gall numbers were intermediate on<br />

hybrids (supp<strong>or</strong>ting the Additive hypothesis). Studies on sawflies on birches (Hanhim/iki et al. 1994), beetles on elms (Hall<br />

and Townsend 1987), and beetles on willows (Soetens et al. 1991, and Fritz unpublished), and various species on oaks<br />

(Aguilar and Boecklen 1992) also indicate that hybrids are often intermediate in suitability f<strong>or</strong> herbiv<strong>or</strong>e growth. Hanhimfiki<br />

et al. (1994) have studied herbiv<strong>or</strong>e perf<strong>or</strong>mance on hybrid and parental birches that grew in a common plantation in Finnish<br />

Lal_land. Of the 14 herbiv<strong>or</strong>e species studied (l 3 sawflies and a moth), I I species had intermediate growth rates on hybrid<br />

leaves compared to parental leaves.<br />

Monogenic and polygenic resistance traits where allele effects on traits are additive will have half the dosage of<br />

alleles, on average, in hybrids compared to parents, unless parents have the same defense mechanisms determined by the<br />

same loci. There is evidence from willows supp<strong>or</strong>ting additive inheritance of phenolic glycosides, tannins, and m<strong>or</strong>phologi-<br />

cal defenses (Meier et al. 1989, Soetens et al. 1991, Nichols-Orians and Fritz, unpublished data). How the altered dosage of<br />

defensive traits influences herbiv<strong>or</strong>es will depend on the sensitivity of the herbiv<strong>or</strong>es to the resistance trait and the presence<br />

of other resistance traits inherited from species that might not n<strong>or</strong>mally be host plants of the herbiv<strong>or</strong>e. A consequence of<br />

additive inheritance of resistance is that backcrossing should result in a reconstitution of defense, rather than further break-<br />

down of resistance, as hypothesized by Whitham (1989) and Paige and Capman (1993). The resistances of backcross<br />

individuals should m<strong>or</strong>e closely resemble that of the recurrent parent as m<strong>or</strong>e parental genes are reinc<strong>or</strong>p<strong>or</strong>ated into the<br />

genome. Whitham et al. (1994) found evidence supp<strong>or</strong>ting <strong>this</strong> hypothesis in their w<strong>or</strong>k on eucalypts. When F1 hybrids<br />

were compared with backcross plants and parents, often backcross plants were intermediate between FI hybrids and parents.<br />

The Dominance hypothesis received supp<strong>or</strong>t from Aculops mites, Phyllocnistis sp. Phyllocolpa spp., and Pontania sp.<br />

in at least one year (Table 1). In <strong>this</strong> hypothesis, herbiv<strong>or</strong>e densities on hybrids resemble those of one parent but differ<br />

significantly from that on the other parental species. Hanhim_iki et al. (1994) found f<strong>or</strong> three species, a moth and two<br />

sawflies, on the Finnish birches that perf<strong>or</strong>mance on hybrids did not differ from that of one parent but did differ from the<br />

other parent, supp<strong>or</strong>ting the Dominance hypothesis. The Dominance hypothesis was proposed by Fritz et al. (1994), and it is<br />

particularly plausible because of the inheritance of defensive chemicals in hybrids of several crop and wild plants. F<strong>or</strong> Lotus<br />

(O'Donoughue et al. 1990), Nicotiana (Huesing et al. 1989), and Papaver (Levy and Milo 1991), chemical defense mecha-<br />

nisms have been shown to be inherited as dominant traits in hybrids. Zangerl and Berenbaum (pers. comm.), in a review of<br />

138


the inheritance patterns of secondary chemicals, found that 13 of 32 compounds had dominant inheritance. Given dominant<br />

inheritance, if the parent possessing the high levels of defensive chemical is resistant, then F1 hybrids should also be resistant.<br />

Paige and Capman (1993) showed that hybrid breakdown in resistance to Pemphigus betae occurs only in backcrosses<br />

of Populus hybrids to the susceptible parent, and F1 hybrids are as resistant as the resistant P. fremomii. This illustrates the<br />

Dominance hypothesis in the cottonwood-aphid system.<br />

There was also considerable supp<strong>or</strong>t f<strong>or</strong> the Hybrid Susceptibility hypothesis. Three herbiv<strong>or</strong>e species consistently<br />

supp<strong>or</strong>ted the Hybrid Susceptibility hypothesis, and over three years 9 of 33 cases supp<strong>or</strong>ted <strong>this</strong> hypothesis. The Hybrid<br />

Susceptibility hypothesis was also supp<strong>or</strong>ted by the analysis of total herbiv<strong>or</strong>e densities in 1993 (Table 1, Fig. 2). Mosseler<br />

(pers. comm.) rep<strong>or</strong>ts that in his garden of F1 hybrids and parents of seven Salix species, hybrids were highly susceptible to<br />

Melamps<strong>or</strong>a rust, but parents were immune. Fritz et al. (1994) has also found f<strong>or</strong> Melamps<strong>or</strong>a rust that hybrids were much<br />

m<strong>or</strong>e susceptible than either parental species. This observation shows that Fl's appear to show breakdown of resistance.<br />

The Hybrid Resistance hypothesis was supp<strong>or</strong>ted only by the Aculops gall mite in 1992. Density of mites was lower<br />

on hybrids and on either parent. Boecklen and Spellenberg (1990) found some evidence f<strong>or</strong> <strong>this</strong> hypothesis in their study of<br />

leaf mining and leaf galling herbiv<strong>or</strong>es on oak. A problem with the supp<strong>or</strong>t f<strong>or</strong> the Hybrid Resistance hypothesis from that<br />

study is that many herbiv<strong>or</strong>e species were combined into feeding guilds f<strong>or</strong> analysis and they did not explicitly test f<strong>or</strong> a<br />

significant departure from the Dominance hypothesis.<br />

The results of <strong>this</strong> study suggest that interspecific hybridization of plants has a variety of effects on resistance of<br />

hybrids and it does not lead only to a pattern of breakdown of resistance (i. e., supp<strong>or</strong>t of the Hybrid Susceptibility hypothesis).<br />

The various herbiv<strong>or</strong>e responses suggest that different resistance traits of plants have different mechanisms of inheritance<br />

and/<strong>or</strong> that different resistance fact<strong>or</strong>s could affect herbiv<strong>or</strong>es differently. An imp<strong>or</strong>tant question is: How can the<br />

hypothesized herbiv<strong>or</strong>e responses be explained by the known patterns of inheritance of resistance traits? The Additive and<br />

Dominance hypotheses both have fairly straightf<strong>or</strong>ward explanations. If resistance is dosage-dependent, then an Additively<br />

inherited trait should lead to supp<strong>or</strong>t f<strong>or</strong> the Additive hypothesis. A dominantly inherited trait would be fully expressed in<br />

hybrids and should result in a pattern of herbiv<strong>or</strong>e resistance that supp<strong>or</strong>ted the Dominance hypothesis. Backcrossing to the<br />

susceptible parent, as seems to have occurred in cottonwoods, could lead to the loss of one <strong>or</strong> m<strong>or</strong>e dominant resistance<br />

genes, thereby resulting in Hybrid Susceptibility of the backcross progeny. Paige and Capman (1993) suggest that there is<br />

m<strong>or</strong>e than one dominant gene involved in resistance to aphids in the cottonwood system.<br />

Additively inherited traits could also result in herbiv<strong>or</strong>e densities that supp<strong>or</strong>t the Hybrid Susceptibility hypothesis.<br />

Herbiv<strong>or</strong>es that require a threshold amount of a chemical defense could have higher densities on hybrids (Hybrid Susceptibility)<br />

if the dosage in hybrids was less than the threshold amount. Hybrid Resistance could result from a mechanism of<br />

dominance inheritance. Hybrids with dominant inheritance of two different resistance traits from each parent would possess<br />

complete expression of two different resistance mechanisms. If particular herbiv<strong>or</strong>es are negatively affected by both resistance<br />

mechanisms, hybrids could be m<strong>or</strong>e resistant than either parent (i.e., Hybrid Resistance). It also seems possible that if<br />

two additively inherited resistance mechanisms interacted synergistically to affect herbiv<strong>or</strong>es Hybrid Resistance could also<br />

result. These alternative mechanisms suggest that various outcomes of interspecific hybridization on resistance are likely and<br />

that mechanisms of inheritance of resistance and their effect on herbiv<strong>or</strong>es need to be investigated in hybrid systems.<br />

The year-to-year variation in supp<strong>or</strong>t f<strong>or</strong> particular hypotheses was an unexpected result of <strong>this</strong> analysis. Seven of<br />

the 11 species had at least one change in which hypothesis was supp<strong>or</strong>ted between adjacent years. These changes could be<br />

considered either: (1) random fluctuations in herbiv<strong>or</strong>e numbers on hybrids and parents <strong>or</strong> accidents of sampling, (2)<br />

consequences of temp<strong>or</strong>arily lower population sizes that made differences difficult to detect, <strong>or</strong> (3) real year-to-year changes<br />

in the expression of resistance of hybrids and/<strong>or</strong> parent species due to variation in environmental fact<strong>or</strong>s. Point 1 is always<br />

possible, but it is impossible to discount using the census data. A number of shifts in herbiv<strong>or</strong>e response occurred between<br />

1991 and 1992 and then back to 1991 patterns in 1993. This may have something to do with the decreased abundance of<br />

herbiv<strong>or</strong>es in 1992 (unpublished results). Lower numbers could have made differences difficult to detect. While point 2 is<br />

possible, it would be wrong to attribute the shifts in supp<strong>or</strong>t f<strong>or</strong> hypotheses to <strong>this</strong> fact<strong>or</strong> alone. Shifts in supp<strong>or</strong>t f<strong>or</strong> hypotheses<br />

may be real changes in the relative expression of resistance between hybrid and parental plants. Environmental variation<br />

(e. g., solar radiation, temperature, nutrients, etc.) is known to affect the expression of resistance among genotypes of many<br />

plant species. It therefbre seems likely that changes in environment could affect resistance of hybrid and parent species of<br />

plants in nature. There were some substantial shifts in the weather among the years of <strong>this</strong> study. The summer of 1992 was<br />

cool and rainy compared to n<strong>or</strong>mal in the n<strong>or</strong>theast, whereas in 1993 there were some long periods of high temperatures and<br />

139


drought (pers. obs.). These fact<strong>or</strong>s might have contributed to the variation in resistance of hybrids. However, experiments<br />

wilt be required to test if tile relative resistance of hybrids shifts with manipulated changes in specific environmental fact<strong>or</strong>s.<br />

Guild membership was not a reliable indicat<strong>or</strong> of similar responses of herbiv<strong>or</strong>es to hybrids and parents. Among the<br />

leaf galling guild, all four hypotheses were supp<strong>or</strong>ted in at least 1 year. Leaf miners supp<strong>or</strong>ted two different hypotheses<br />

(Additive and Hybrid Susceptibility) considering the 3 years together (Table 2). These results indicate that combining the<br />

densities of two <strong>or</strong> m<strong>or</strong>e species in the same guild f<strong>or</strong> hybrid-parent comparisons will obscure the underlying patterns of<br />

herbiv<strong>or</strong>e densities on hybrids and parents. In <strong>this</strong> analysis, numbers of Phyllocolpa nigrita and Phyllocolpa sp. nov. were<br />

combined in 1992 and 1993, because densities of these species had not been distinguished on S. sericea of the willow species<br />

in 1991. Theref<strong>or</strong>e, the conclusions drawn from these species should be considered with caution.<br />

SUMMARY<br />

We studied herbiv<strong>or</strong>y of two species of willows (Salix sericea and S. eriocephala) and their interspecific hybrids to<br />

test four alternative hypotheses concerning the effects of hybridization on plant resistance. Individually marked plants were<br />

identified using m<strong>or</strong>phological traits in the field, and RAPD band analysis was used to verify the genetic status of some<br />

parental and hybrid plants. The densities of 11herbiv<strong>or</strong>e species were compared between 2 parents and their hybrids in the<br />

field. We found the most supp<strong>or</strong>t f<strong>or</strong> the Additive hypothesis and the Hybrid Susceptibility hypothesis over the three years.<br />

We found some evidence f<strong>or</strong> the Dominance hypothesis, and one species in one year supp<strong>or</strong>ted the Hybrid Resistance<br />

hypothesis. Guild membership was not a good predict<strong>or</strong> of similar responses of species to hybrid versus parental plants.<br />

This study demonstrates the diversity of responses of phytophages in response to interspecific hybridization, and indicates the<br />

presence of year-to-year variation, which might indicate the influence of environmental variation on the expression of hybrid<br />

resistance.<br />

ACKNOWLEDGMENTS<br />

We would like to thank Len and Ellie Sosnowski f<strong>or</strong> allowing us to w<strong>or</strong>k on their property. We thank S. Kaufman,<br />

N. Murphy, P. Zee, M. Momot, A. Samsel, A. Dao, J. Hama, A. Samsel, and G. Sylvan f<strong>or</strong> help in the field. This research<br />

was supp<strong>or</strong>ted by the National Science Foundation grants BSR 89-17752 to R. S. Fritz and DEB 92-07363 to R. S. Fritz and<br />

C. M. Orians, and by the Vassar College General <strong>Research</strong> Fund.<br />

LITERATURE CITED<br />

AGUILAR, J.M. and BOECKLEN, W.J. 1992. Patterns of herbiv<strong>or</strong>y in the Quercus grisea x Quercus gambelii species<br />

complex. Oikos 64: 498-504.<br />

ARGUS, G.W. 1974. An experimental study of hybridization and pollination in Salix (willows). Can. J. Bot. 52: 1613-1619.<br />

ARGUS, G.W. 1986. The genus Salix (Salicaceae) in the southeastern United States. Syst. Bot. Monogr. 9: 1-170.<br />

BOECKLEN, W.J.and SPELLENBERG, R. 1990. Structure of herbiv<strong>or</strong>e communities in two oak (Quercus spp.) hybrid<br />

zones. Oecologia 85: 92-100.<br />

DRAKE, D.W. 1981. Reproductive success of two Eucalyptus hybrid populations. II. Comparison of predispersal seed<br />

parameters. Aust. J. Bot. 29: 37-48.<br />

ERICSON, L., BURDON, J.J., and WENNSTROM, A. 1993. Inter-specific host hybrids and phalacrid beetles implicated in<br />

the local survival of smut pathogens. Oikos 68: 393-400.<br />

FLOATE, K.D., KEARSLEY, M.J.C., and WHITHAM, T.G. 1993. Elevated herbiv<strong>or</strong>y in plant hybrid zones: Chrysomela<br />

cot_uens, Populus and phenological sinks. Ecology 74: 2056-2065.<br />

140


FLOATE, K.D. and WHITHAM, T.G. 1993. The "hybrid bridge" hypothesis: host shifting via plant hybrid swarms. Am.<br />

Nat. 141 : 651-662.<br />

FRITZ, R.S., NICHOLS-ORIANS, C.M., and BRUNSFELD, S.J. 1994. Interspecific hybridization of plants and resistance<br />

to herbiv<strong>or</strong>es: Hypotheses, genetics, and variable responses in a diverse community. Oecologia 97:106-117.<br />

HALL, R.W. and TOWNSEND, A.M. 1987. Suitability of Ulmus wilsoniana, the 'Urban' elm, and their hybrids f<strong>or</strong> the ehn<br />

leaf beetle, Xa_lthogaleruca luteola (Mtiller) (Coleoptera: Chrysomelidae). Environ. Entomol. 16: 1042-1044.<br />

HANHIM_.KI, S., SENN, J., and HAUKIOJA, E. 1994. Perf<strong>or</strong>mance of insect herbiv<strong>or</strong>es on hybridising trees: the case of<br />

the subarctic birches. J. Anita. Ecol. in press.<br />

HUESING, J., JONES, D., DEVEERNA, J., MYERS, J., COLLINS, G., SEVERSON, R., and SISSON, V. 1989. Biochemical<br />

investigations of antibiosis material in leaf exudate of wild Nicotiana species and interspecific hybrids. J. Chem.<br />

Ecol. 15: 1203-1217.<br />

KEIM, R, PAIGE, K.N., WHITHAM, T.G., and LARK, K.G. 1989. Genetic analysis of an interspecific swarm of Populus:<br />

Occurrence of unidirectional introgression. Genetics 123: 557-565.<br />

LEVY, A. and MILO, J. 1991. Inheritance of m<strong>or</strong>phological and chemical characters in interspecific hybrids between<br />

Pat)aver bracteatum and Papaverpseudo-<strong>or</strong>ientale. The<strong>or</strong>. Appl. Genet. 81: 537-540.<br />

MEIER, B., BETTSCHART, A., SHAO, Y., and LAUTENSCHLAGER, E. 1989. Einsatz der modernen HPLC f<strong>or</strong><br />

chemotaxonomische Utersuchungen m<strong>or</strong>phologisch schwer zu differenzierender Salix-Hybriden. Planta Medica 55:<br />

213-214.<br />

MOSSELER, A. and PAPADOPOL, C.S. 1989. Seasonal isolation as a reproductive barrier among sympatric Salix species.<br />

Can. J. Bot. 67: 2563-2570.<br />

O'DONOUGHUE, L.S., RAELSON, J.V., and GRANT, W.E 1990. A m<strong>or</strong>phological study of interspecific hybrids in the<br />

genus Lotus (Fabaceae). Can. J. Bot. 68: 803-812.<br />

PAIGE, K.N., CAPMAN, W.C., and JENNETTEN, R 1991. Mitochondrial inheritance patterns across a cottonwood hybrid<br />

zone: Cytonuclear disequilibria and hybrid zone dynamics. Evolution 45: 1360-1369.<br />

PAIGE, K.N. and CAPMAN, W.C. 1993. The effects of host-plant genotype, hybridization and environment on gall aphid<br />

attack and survival in cottonwood: The imp<strong>or</strong>tance of genetic studies and the utility of RFLP's. Evolution 47: 36-45.<br />

RICE, W.R. 1989. Analyzing tables of statistical tests. Evolution 43: 223-225.<br />

SAS INSTITUTE. 1985. SAS user's guide: Statistics, Version 5 Edition. SAS Institute, Inc. Cary, NC.<br />

SOETENS, R, ROWELL-RAHIER, M., and PASTEELS, J.M. 1991. Influence of phenolglucosides and trichome density<br />

on the distribution of insects herbiv<strong>or</strong>es on willows. Entomol. Exp. Appl. 59: 175-187.<br />

WHITHAM, T.G. 1989. Plant hybrid zones as sinks f<strong>or</strong> pests. Science 244: 1490-1493.<br />

WHITHAM, T.G., MORROW, RA., and POTTS, B.M. 1991. Conservation of hybrid plants. Science 254: 779-780.<br />

WHITHAM, T.G., MORROW, RA., and POTTS, B M. 1994. Plant hybrid zones as centers of biodiversity: The herbiv<strong>or</strong>e<br />

community of two endemic Tasmanian eucalypts. Oecologia, in press.<br />

141


F1 HYBRID SPRUCES INHERIT THE PHYTOPHAGOUS INSECTS<br />

OF THEIR PARENTS<br />

WILLIAM J. MATTSON, ROBERT K. LAWRENCE, and BRUCE A° BIRR<br />

<strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong><br />

B-7 Pesticide <strong>Research</strong> Center, Michigan State University, East Lansing, MI 48824, USA<br />

INTRODUCTION<br />

Interspecific hybridization in plants is a widespread and imp<strong>or</strong>tant natural phenomenon. In fact, Stace (1987)<br />

estimates that at least 50% of the extant angiosperm species have been derived via hybridization. Natural hybrids are also<br />

common among the gymnosperms (Wright 1955, Hanover and Wilkinson 1969, Schmidt-Vogt 1977, Zobel and Talbert 1984).<br />

In spite of the ecological and evolutionary imp<strong>or</strong>tance of plant hybrid zones (Remington 1968, Barton and Hewitt 1985),<br />

little is yet known about how they affect the dynamics of plant/herbiv<strong>or</strong>e relationships. Whitham (1989) speculated that<br />

hybrids may be particularly prone to herbiv<strong>or</strong>y because usual resistance mechanisms derived from the parents are po<strong>or</strong>ly<br />

expressed. F<strong>or</strong> example, polygenic resistance traits might be intermediate in F1 hybrids, whereas monogenic and oligogenic<br />

traits might be fully expressed <strong>or</strong> not at all depending on the mode of gene action (e.g., dominant <strong>or</strong> recessive, etc.), and on<br />

whether there is genetic dysfunction preventing n<strong>or</strong>mal gene expression (Fritz et al. 1994, Whitham et al. 1994). Furtherm<strong>or</strong>e,<br />

to compound these problems, hybrids may attract specialist phytophages from both parent species, thereby being<br />

predisposed to a much larger pool of potential consumers than either of the pure parent types (Whitham et al. 1994).<br />

Insect Loading in Relation to a Plant's A/R Vect<strong>or</strong>s and S/D Vect<strong>or</strong>s<br />

To examine the issue of hybrid resistance to insects in a m<strong>or</strong>e mechanistic framew<strong>or</strong>k, we consider that there are at<br />

least two sets of plant traits affecting insect loading on plants: (1) the mix of properties which elicit insect attraction/<br />

arrestmenffrepulsion (e.g., m<strong>or</strong>phology, electromagnetic and biochemical cues, abiotic requirements, etc.) and thus govern the<br />

number of insect species that attempt to colonize it, and (2) the mix of properties that are sustaining/defensive f<strong>or</strong> each of the<br />

would be colonizers as they attempt to feed, oviposit, and reproduce. These two sets of properties, which together set the<br />

potential f<strong>or</strong> insect loading (total insect biomass <strong>or</strong> numbers), can be represented as two vect<strong>or</strong>s: the species attraction/<br />

repulsion (A/R) vect<strong>or</strong>, [a_,a2,a3..an],and the suitability/defense (S/D) vect<strong>or</strong> [Sl,S2,S3..s]. Each species of insect, Si, is<br />

represented by an element (a)in the species A/R vect<strong>or</strong>, having a numerical value (0-1) that is prop<strong>or</strong>tional to the plant's<br />

average capacity to attract and arrest individuals of the particular insect species. Likewise, the S/D vect<strong>or</strong> contains specific<br />

elements (s) that c<strong>or</strong>respond to and represent the plant's average suitability, ranging from zero, where an insect colonizes a<br />

plant but cannot complete development on it (immunity), to 1 where the plant fulfills all of the insect's requirements, f<strong>or</strong> each<br />

species in the species A/R vect<strong>or</strong>. The product of the two vect<strong>or</strong>s, [ayL+a2s2+..as ], represents the plant's overall potential<br />

"resistance expression", <strong>or</strong> potential insect loading per plant.<br />

To give an extreme example f<strong>or</strong> one insect: the white pine weevil, Pissodes strobi, is highly attracted to Picea glauca<br />

shoots as evidenced by the great number of feeding scars it produces (i.e., a = 1), but it can only rarely reproduce in these<br />

shoots (i.e., s = 0), with the consequence that P glauca is considered "resistant" to the weevil (Lanier 1983). Actual insect<br />

loading depends, in addition, on the quantity and genetic diversity of insects (N) that continually attempt colonization. The<br />

total number of non-zero elements (all those a_> 0) in an individual plant's A/R <strong>or</strong> species vect<strong>or</strong> (its species richness <strong>or</strong><br />

species loading) is positively linked to the size and density of its population, the total areal distribution of the species, and<br />

negatively to its biochemical uniqueness wherein it occurs (f<strong>or</strong> a review, see Tahvanainen and Niemela 1987). F<strong>or</strong> many<br />

natural hybrids, their populations and geographical distributions are usually small (compared to their parents) thereby<br />

exposing them to fewer potential colonists, but their low biochemical uniqueness <strong>or</strong> high similarity to the two nearby parent<br />

species will likely have a counteracting influence making them highly susceptible to the reservoirs of consumers fiom two<br />

plant species (Whitham et al. 1994).<br />

Mattson, W.J., Niemel_, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

142<br />

;fl ¸_


Elements of the Suitability/Defense Vect<strong>or</strong> in Hybrids: Expression of S/D Traits<br />

Most w<strong>or</strong>k on resistance expression (i.e., a_s_)in natural and artificial interspecies hybrids has focused on suitability/<br />

defense (s) expression f<strong>or</strong> particular insects, assuming that a plant's attractiveness (a) is high and invariant. Fritz et al.<br />

(1994) hypothesized four expression scenarios: (1) the "additive" case where the hybrid's S/D ranking (s_) against an insect<br />

is the simple mean of its parents, p_, and p: (sh_= (s_ + s2)/2), (2) the "dominance" case where the hybrid is consistently like<br />

one of the parents, e.g., p_(sh_= s_), (3) the "susceptibility" case where the hybrid has m<strong>or</strong>e insects than either of the parents<br />

(s_a< sh_> s2), and (4) the "resistance" case where the hybrid has fewer insects than either parent (s_ > sh_< s:).<br />

Most studies with trees have shown that hybrid S/D expression (s) against an insect is not cleanly predictable but<br />

varies with the particular insect and the host plant species (Boecklen and Larson 1994, Fritz et al. 1994). F<strong>or</strong> example,<br />

Manley and Fowler (11969)and Osawa (1989) discovered that the severity of spruce budw<strong>or</strong>m, Ch<strong>or</strong>istonenrafumiferana,<br />

defoliation of spruce trees was linked to the direction of their introgression between black and red spruces, P. maria_za and P.<br />

rubens. Hybrids closer to pure red were severely defoliated as were pure reds, whereas hybrids closer to pure black were<br />

only lightly defoliated as were pure blacks. Several hybrid elms, Ulmus, were consistently intermediate in resistance to the<br />

elm leafbeetle, Xanthogaleruca luteola, whenever one parent species was susceptible and the other was resistant (Hall and<br />

Townsend 1987, Hall et al. t987). Hybrids of P. sitchensis x P. glauca are usually resistant to Pissodes strobi, and so are<br />

such backcrosses to white spruce, the resistant parent whereas sitka spruce is highly susceptible (Mitchell et al. 1974, Copes<br />

and Beckwith 1977). Similarly, the Pinus coulteri x P jeffreyi hybrid carries resistance against the weevil, Cylirzrocopturus<br />

eatoni, that the coulteri parent possesses, but thejeffreyi parent does not (Miller 1950). Hybrids of several species of<br />

Japanese pines, Pinus, all apparently inherited their "resistance" against the scale, Matscoccus matsumurae, from the m<strong>or</strong>e<br />

susceptible parent species (McClure 1985).<br />

M<strong>or</strong>e recently, Floate et al. (1993) rep<strong>or</strong>ted that Chrysomela confluens was invariably m<strong>or</strong>e abundant in a hybrid<br />

swarm of Populusfremontii x P angustifolia than in the parental populations. Whitham et al. (1994) rep<strong>or</strong>ted that of some<br />

40 phytophagous taxa in a hybrid swarm of Eucalyptus risdonii x E. amygdalina only two were m<strong>or</strong>e abundant on FI hybrids<br />

than on the parents. Specialist insects were most abundant on hybrid backcrosses to their fav<strong>or</strong>ed parent species, and<br />

generalist insects were generally m<strong>or</strong>e abundant on all hybrid types. Hanhimaki et al. (1994) rep<strong>or</strong>ted that in the case of 14<br />

insects tested on hybrids of Betula pubescens ssp. t<strong>or</strong>tuosa x B. nana, the hybrids were consistently equivalent to B.<br />

pubescens, but superi<strong>or</strong> to B. nana. Fritz et al. (1994) tested 11 insects and one leaf rust on hybrids of Salix sericea x S.<br />

eriocephala, and concluded that 3 supp<strong>or</strong>ted the additive, 2 the dominance, and 7 the susceptibility expression hypotheses.<br />

Boecklen and Larson (1994) measured densities of 8 species of galling wasps on hybrids of Quercus grisea x Q. gambelii,<br />

and rep<strong>or</strong>ted that 3 supp<strong>or</strong>ted the additive, 2 the dominance, 2 the susceptibility, and 1the resistance expression hypotheses.<br />

In the case of non-woody species, plant breeders have created innumerable interspecies hybrids which have become<br />

the c<strong>or</strong>nerstone of breeding programs f<strong>or</strong> the development of insect resistance in food and f<strong>or</strong>age plants. Consequently, most<br />

resistance traits useful against native insects and pathogens have been derived from crosses to related but allopatric plant<br />

species (Harris 1975, 1980 1982; Bailey 1983). These studies reveal a surprisingly large number of cases f<strong>or</strong> which insect<br />

resistance seems to be either monogenic <strong>or</strong> oligogenic, but polygenic inheritance is also imp<strong>or</strong>tant (Hanover 1980, Harris<br />

1982, Zobel and Talbert 1984, Singh 1986, Carson and Carson 1989, Geiger and Heun 1989). M<strong>or</strong>eover, the resistance genes<br />

are usually, but not always dominant in their expression (Harris 1982, Singh 1986), and evidence suggests that gene linkages<br />

and multiple resistance alleles are uncommon. Resistance genes appear to be insect specific, not conferring cross protection<br />

against many species (Harris 1982).<br />

Elements of the Species A/R Vect<strong>or</strong> in Hybrids: Expression of A/R Traits<br />

General plant traits (e.g., crown f<strong>or</strong>m, height, growth rate, phenology, etc.) in hybrids are often intermediate between<br />

those of its parents (implying that S,_= (a_ + a2)/2) (Hanover and Wilkinson 1969, Bongarten and Hanover t982, Yeh and<br />

Arnott 1986) which suggests that interspecies hybrids will probably have the unique gestalt of properties to attract the all the<br />

specialist insects from both parents (Miller and Strickler 1984). In t_ct, Whitham et al. (1994) and M<strong>or</strong>row et al. (1994) have<br />

found that <strong>this</strong> is true in a eucalypt hybrid swarm. They rep<strong>or</strong>ted that the average hybrid tree supp<strong>or</strong>ted 53% m<strong>or</strong>e phytophagous<br />

insect and fungal species than equivalent trees from the pure parental stands.<br />

143


Objectives<br />

Given the imp<strong>or</strong>tance of a plant's herbiv<strong>or</strong>e loading, i.e., the number of herbiv<strong>or</strong>e species <strong>or</strong> non-zero elements in its<br />

species vect<strong>or</strong>, <strong>this</strong> study tested the hypothesis that artificial white x blue spruce FI hybrids will inherit the phytophages from<br />

each of their parent tree species, with the consequence that the number of elements in the hybrid species vect<strong>or</strong> will be the<br />

sum of the unique (specialists) plus the shared <strong>or</strong> common (generalists) elements of both parents. Because hybrid spruces do<br />

not occur naturally in Michigan, we only conducted tests in several test plantations where the trees are exposed to the natural<br />

phytophages of white spruce but not those of the other parent, blue spruce, a western N<strong>or</strong>th American tree species. However,<br />

one western spruce galling insect, Adelges cooleyi, is common in eastern N<strong>or</strong>th America on <strong>or</strong>namental blue spruce. We also<br />

tested whether total insect loading on hybrids, i.e., the total number of individuals (pooled over all phytophage species) found<br />

per unit of foliage per plant, is equal to <strong>or</strong> greater than that found on each of the parent species.<br />

METHODS<br />

Hybrid spruces (P glauca x P. pungens) were sampled in 1991 and 1992 at two different locations in Michigan, along<br />

with populations of the parent species in <strong>or</strong>der to assess differences in species, and total insect loading. One sample site<br />

occurred in south-central Michigan at the Michigan State University Kellogg Experimental F<strong>or</strong>est. The other occurred some<br />

450 km to the n<strong>or</strong>th in Michigan's Upper Peninsula at the Michigan State University Dunbar Experimental F<strong>or</strong>est. Kellogg<br />

(n = 40) and Dunbar (n = 24) sample trees were all F1 hybrids. Equivalent numbers of parent spruces were likewise sampled<br />

from adjacent <strong>or</strong> nearby (< 1 kin) plantations.<br />

At each site, trees were randomly selected, and examined twice per growing season (late May-early June, and late<br />

June-mid July). Trees were first examined visually to sc<strong>or</strong>e f<strong>or</strong> adelgid galls (Adelges abietis, A. cooleyi, and Pineus similis).<br />

Next they were sampled using standard beating methods whereby the apical half (45 cm length) of a midcrown branch on the<br />

west side was held over a specially designed stainless steel collection pan, and the branch was gently tapped with a wooden<br />

dowel f<strong>or</strong> I0 seconds to encourage free feeding species to drop off into the pan. Such insects were gathered in vials and then<br />

st<strong>or</strong>ed in a freezer f<strong>or</strong> later identification and counting.<br />

Data from both June and July samples (numbers of insect species, and total insect counts per tree) were analyzed<br />

separately using a completely randomized ANOVA, after log (X + 0.1) transf<strong>or</strong>mation. Means f<strong>or</strong> each tree species were<br />

ranked and then separated using the SNK multiple range test.<br />

RESULTS<br />

Gall-f<strong>or</strong>ming Adelgid Specialists from Both Parent Spruces Successfully Attack Hybrids<br />

At the Kellogg experimental tbrest in southern Michigan, there were significant populations of both Adelges abietis,<br />

and A. cooleyi, largely specialists, respectively, on P. glauca, and P pungens and their near relatives (Furniss and Carolin<br />

1977). We found that the ihybrids contained substantial numbers of both kinds of adelgids, whereas the parent spruces had<br />

primarily their respective specialist adelgids (Table 1). One glauca individual had evidence of two po<strong>or</strong>ly f<strong>or</strong>med A. cooleyi<br />

galls, but in general glauca in eastern N<strong>or</strong>th America are not susceptible to the eastern A. cooleyi populations (personal<br />

observations) which live largely on <strong>or</strong>namental blue and Engelmann spruces introduced from western N<strong>or</strong>th America.<br />

Similarly, there was only min<strong>or</strong> evidence f<strong>or</strong> A. abietis on pungens: one tree, with one small, po<strong>or</strong>ly f<strong>or</strong>med gall. Although<br />

Table 1 shows very low populations ofA. cooleyi on pungens (one infested tree in a sample of 30-42), <strong>this</strong> was because the<br />

adelgid population had drastically declined in the preceding 2 years (1989, 1990) in the sampled pungens plantation which<br />

was about 1 kln from the hybrids. In 1988 and earlier there were much larger populations of A.cooleyi there: 76 of 100<br />

sample trees were infested (Mattson, unpublished data). Pineus simitis, another galling adelgid that is largely a specialist of<br />

glauca and near relatives, was not abundant enough to test hybrid susceptibility.<br />

At the Dunbar site in n<strong>or</strong>thern Michigan, A. cooleyi and P. similis populations were negligible, so we could not<br />

measure their colonization of the spruce hybrids. But, we did find there that A. abietis readily attacked hybrids, but not so<br />

readily pungens, as shown in the following tabulation of the percentage of trees attacked: white (84), hybrids (83), blue (15).<br />

144


Table l.--Comparing the mean numbers of insects and adelgid galls on blue, hybrid, and white spruces, using visual<br />

examination of two random midcrown branches per sample tree at the Kellogg Experimental F<strong>or</strong>est during late<br />

May-early June, and mid-late July. N = 30 - 42 trees.<br />

Insect Species Time period l: nos.! 2 branches Time period 2: nos./2 branches<br />

B.Spruce Hybrids W.Spruce B.Spruce Hybrids W.Spruce<br />

A. abietis galls 0.07b _ 1.14a 2.25a 0.00b 2.20a 2.27a<br />

A. cooleyi galls 0.03b 0.36a 0.00b 0.00b 0.33a 0.07b<br />

A. cooleyi alates 0.17b 24.76a 0.00b 0.00b 4.66a 0.00b<br />

Pineus similis galls O.OOa O.OOa O.19a O.OOa O.OOa O.OOa<br />

Means with different letters within time periods are significantly (p < 0.05) different.<br />

Species Loading, and Total Insect Loading Comparable Among Hybrids and White Spruce<br />

At Kellogg, the hybrid and white spruce herbiv<strong>or</strong>e species loadings were statistically equivalent (ca. 3 - 3.3 per<br />

branch) and significantly higher than those of blue (I.9) in the first sample, but not in the second, when all three spruces were<br />

statistically equivalent. At Dunbar, the hybrids, and white spruce also had equivalent herbiv<strong>or</strong>e loadings (ca. 3 - 3.8 species<br />

per branch) which were significantly higher than those (ca. 1.2 - 1.5 per branch) of blue spruce in both first and second<br />

samples (Table 2).<br />

Table 2.--Comparing mean numbers of phytophagous insect species, and total insect numbers occurring in beating samples<br />

from blue, hybrid, and white spruces at two study sites in Michigan.<br />

Study area and Mean nos. insect species per branch Mean nos. of insects per branch<br />

sample period<br />

B.Spruce Hybrids W.Spruce B.Spruce Hybrids W.Spruce<br />

Dunbar- 1 1.50b _ 2.96a 3.79a 2.42b 4.17a 6.71 a<br />

Dunbar-2 l. 17b 3.21 a 3.00a 1.25b 4.33a 4.71 a<br />

Kellogg-1 1.87b 3.31a 2.97a 1.13b 4.27a 4.23a<br />

Kellogg-2 2.97a 2.17a 2.90a 2.20a 2.12a 1.97a<br />

Means having different letters within an area and sample period categ<strong>or</strong>y are significantly (p < 0.05) different.<br />

At Kellogg, there were 20 species of phytophagous insects on the hybrids, of which they shared 12 in common (60%)<br />

with white spruce (Fig. l). At Dunbar, we found twice as many (41 ) species of phytophagous insects on the hybrids, of<br />

which 33 (80.5%) they shared in common with white spruce (Fig. I). About 12-15% of the hybrid's phytophagous insects<br />

were "unique" at both sites, because they did not share these species with either white <strong>or</strong> blue spruce. About 49-60% of their<br />

total species pool, they shared with blue spruce. These data confirm the expectation that the phytophagous insects occurring<br />

on the hybrids are largely those commonly associated with white spruce. The same is true f<strong>or</strong> blue spruce, which had a total<br />

of 23 species at Kellogg and 24 at Dunbar, of which 19 were shared in common with white spruce at both sites.<br />

145


Blue White White<br />

2 5% Blue 0%<br />

7.:<br />

Jnique<br />

" "_ "_''''''"'- Unique 12.2%<br />

W&B 15% W&B<br />

35% 41.5%<br />

Figure l.--The percent distribution of the phytophagous insect fauna of hybrid spruces that are shared with blue spruce only,<br />

white spruce only, both white and blue spruce, and unshared <strong>or</strong> unique at two sites: Kellogg (left side) in southern,<br />

and Dunbar (right side) in n<strong>or</strong>thern Michigan. At Kellogg there were 20 and at Dunbar there were 41 species of<br />

insects found in beating samples from the hybrids.<br />

DISCUSSION<br />

The data clearly suggest that white x blue spruce FI hybrids inherit most of the phytophagous insect species of their<br />

white spruce parents. Although we cannot claim that they similarly inherit most of the phytophagous insect species of their<br />

blue spruce parentage because the hybrids were not exposed to these insects in Michigan, we can never-the-less say that they<br />

readily inherited at least one of them, A. cooleyi. Likewise, total insect loading on the hybrids was comparable to that of<br />

white spruce.<br />

Thus, the data do not falsify the hypothesis that the number of phytophages (i.e., the number of non-zero elements) in<br />

the hybrid species vect<strong>or</strong> will be the sum of the unique and shared species of its parents. We fully expect that if the hybrids<br />

were growing adjacent to large, natural populations of blue spruce, they would readily inherit its typical phytophages.<br />

Whitham et al. (1994) emphasized <strong>this</strong> very imp<strong>or</strong>tant point f<strong>or</strong> Tasmanian eucalyptus hybrids, rep<strong>or</strong>ting that the average<br />

hybrid had 53% m<strong>or</strong>e phytophagous species than equivalent trees in pure parental stands.<br />

Neither do our data falsify the hypothesis that the total insect loading (i.e., the total number of all individuals of all<br />

insect species per unit foliage) on hybrids will be equal to <strong>or</strong> greater than the average loading of each of its parents. This may<br />

be true because the defensive traits (sh_)fbr any insect in the hybrids will probably never be m<strong>or</strong>e strongly expressed than in<br />

the less suitable <strong>or</strong> better defended of the two parent species (e.g., if sL_> s2_' then s2_> sh_), and in many if not most cases, it<br />

will be less po<strong>or</strong>ly expressed (e.g., the additive, the susceptibility, and the dominant (high defense is recessive) expression<br />

scenarios of Fritz et al. 1994). At the level of the whole phytophagous insect community, the hybrid studies of Boecklen and<br />

Larson (1994), Fritz et al. (1994), Hanhimaki et al. (1994), and Whitham et al. (1994) unequivocally supp<strong>or</strong>t <strong>this</strong> proposition.<br />

Hybrids may theref<strong>or</strong>e generally sustain m<strong>or</strong>e herbiv<strong>or</strong>es and herbiv<strong>or</strong>e injury than either pure parent species.<br />

SUMMARY<br />

FI hybrids of the spruces Picea glauca x Picea pungens were studied at two Michigan locations to measure and<br />

compare their total phytophagous insect fauna (i.e., species and total loading) with that of the parent species in adjacent <strong>or</strong><br />

nearby plantations. Hybrids had species compositions, and total insect loading nearly identical with that of Picea glauca. In<br />

addition, at least one adelgid specialist from P. pungens readily colonized and successfully reproduced on the hybrids. These<br />

data supp<strong>or</strong>t the hypotheses that hybrids will (a) inherit the insect specialists from both parents, and (b) have an overall insect<br />

loading that equals <strong>or</strong> exceeds that of each parent.<br />

146<br />

i:i_ ¸J'"


ACKNOWLEDGEMENTS<br />

The auth<strong>or</strong>s sincerely thank Drs. J. Kouki and J. Spence f<strong>or</strong> their discussions about and their helpful critiques of<br />

earlier drafts of <strong>this</strong> manuscript. They also thank the personnel of the W.K. Kellogg F<strong>or</strong>est f<strong>or</strong> their unstinting assistance.<br />

LITERATURE CITED<br />

BAILEY, J.A. 1983. Biological perspectives of host-pathogen interactions, p. 1-32. In Bailey, J.A. and Deverall, B.J., eds.<br />

The Dynamics of Host Defense. Academic Press, N.Y.<br />

BAILEY, J.A. 1980. Arthropod-plant interactions related to agriculture, emphasizing host plant resistance, p. 23-51. In<br />

Harris, M.K., ed. Biology and Breeding f<strong>or</strong> Resistance to Arthropods and Pathogens in Agricultural Plants. Texas<br />

A&M Univ., College Sta, TX.<br />

BAILEY, J.A. 1982. Genes f<strong>or</strong> resistance to insects, emphasizing host-parasite interactions in Carya, p. 72-83. In<br />

Heybroek, H.M., tephan, B.R., and von Weissenberg, K., eds. Resistance to Diseases and Pests in F<strong>or</strong>est Trees.<br />

Pudoc, Wageningen, Netherlands.<br />

BARTON, N.H. and G.M. Hewitt, G.M. 1985. Analysis of hybrid zones. Ann. Rev. Ecol. Syst. 16: 1113-148.<br />

BOECKLEN, W.J. and SPELLENBERG, R. 1990. Structure of herbiv<strong>or</strong>e communities in two oak (Quercus spp.) hybrid<br />

zones. Oecologia 85: 92-100.<br />

BOECKLEN, W.J. and LARSON, K.C. 1994. Gall-f<strong>or</strong>ming wasps (Hymenoptera: Cynipidae) in an oak hybrid zone: testing<br />

hypotheses about hybrid susceptibility to herbiv<strong>or</strong>es, p. 110-120. In Price, P., Mattson, W.J., and Baranchikov, Y.N.,<br />

eds. The ecology and evolution of gall-f<strong>or</strong>ming insects. Gen. Tech. Rep. NC-174. St. Paul, MN: U.S. Department of<br />

Agriculture, F<strong>or</strong>est Service.<br />

BONGARTEN, B.C. and J.W. HANOVER 1982. Hybridization among white, red, blue, and white x blue spruces. F<strong>or</strong>. Sci.<br />

28: 129-134.<br />

CARSON, S.D. and CARSON, M.J. 1989. Breeding f<strong>or</strong> resistance in f<strong>or</strong>est trees--a quantitative genetic approach. Ann.<br />

Rev. Phytopath. 27: 373-395.<br />

COPES, D.L. and BECKWITH, R.C. 1977. Isoenzyme identification of Picea glauca, P sitchensis, and P. lutzii populations.<br />

Bot. Gaz. 138: 512-521.<br />

FLOATE, K.D., KEARSLEY, M.J.C., and WHITHAM, T.G. 1993. Elevated herbiv<strong>or</strong>y in plant hybrid zones: Chrysomela<br />

confluens, Populus and phenological sinks. Ecology 74: 2056-2065.<br />

FRITZ, R.S., NICHOLS-ORIANS, C.M, and BRUNSFELD, S.J. 1994. Interspecific hybridization of plants and resistance<br />

to herbiv<strong>or</strong>es: hypotheses, genetics, and variable responses in a diverse herbiv<strong>or</strong>e community. Oecologia 97: 106-<br />

117.<br />

FURNISS, R.L. and CAROLIN, V.M. 1977. Western f<strong>or</strong>est insects. Misc. Publ. 1339. Washington, DC: U.S. Department<br />

of Agriculture, F<strong>or</strong>est Service. 654 p.<br />

GEIGER, H.H. and HEUN, M. 1989. Genetics of quantitative resistance to fungal diseases. Ann. Rev. Phytopath. 27: 317-<br />

341.<br />

HALL, R.W., TOWNSEND, A.M., and BARGER, J.H. 1987. Suitability of thirteen different host species f<strong>or</strong> elm leaf<br />

beetle, Xanthogaleruca luteola (Coleoptera: Chrysomelidae). J. Environ. H<strong>or</strong>t. 5: 143-145.<br />

147


HALL, R.W. and TOWNSEND, A.M. 1987. Suitability of Ulmus wilsoniana, the 'urban' elm, and their hybrids fbr the elm<br />

leaf beetle, Xanthogaleruca luteola (Muller)(Coleoptera: Chrysomelidae). Environ. Ent. 16:1042-1044.<br />

HANHIMAKI, S., SENN, J., and HAUKIOJA, E. 1994. Perf<strong>or</strong>mance of insect herbiv<strong>or</strong>es on hybridizing trees: the case of<br />

the subarctic birches. J. Animal Ecol. 63: 163-175.<br />

HANOVER, J.W. and WILKINSON, R.C. 1969. A new hybrid between blue spruce and white spruce. Can. J. Bot. 47:<br />

1693-1700.<br />

HANOVER, J.W. 1980. Breeding f<strong>or</strong>est trees resistant to insects, p. 487-511. In Maxwell, F.G. and Jennings, ER., eds.<br />

Breeding Plants Resistant to Insects. John Wiley and Sons, NY.<br />

HARRIS, M.K. 1975. Allopatric resistance: searching f<strong>or</strong> sources of insect resistance f<strong>or</strong> use in agriculture. Environ. Ent. 4:<br />

661-669.<br />

KUDRAY, G.M. and HANOVER, J.W. 1980. A preliminary evaluation of the Spartan spruce (Picea glauca x P. pungens<br />

hybrid). Mich. State Univ. Agric.<br />

LANIER, G. 1983. Integration of visual stimuli, host od<strong>or</strong>ants, and pheromones by bark beetles and weevils in locating and<br />

colonizing trees, p. 161-172. In Ahmad, S., ed. Herbiv<strong>or</strong>ous insects: host seeking behavi<strong>or</strong> and mechanisms. Acad.<br />

Press, NY. 257 p.<br />

MANLEY, S.A.M. and FOWLER, D.E 1969. Spruce budw<strong>or</strong>m defoliation in relation to introgression in red and black<br />

spruce. F<strong>or</strong>. Sci. 15: 365-366.<br />

MCCLURE, M.S. 1985. Susceptibility of pure and hybrid stands of Pinus to attack by Matsucoccus matsumurae in Japan<br />

(Homoptera: Coccoidea: Margarodidae). Environ. Ent. 14: 535-538.<br />

MILLER, J.R. and STRICKLER, K.L. 1984. Finding and accepting host plants, p. 127-157. In Bell, W.J. and Carde, R.T.,<br />

eds. Chemical Ecology of Insects. Sinauer Assoc., Sunderland, MA.<br />

MILLER, J.M. 1950. Resistance of pine hybrids to the pine reproduction weevil. F<strong>or</strong>. Res. Notes 68. U.S. Department of<br />

Agriculture, F<strong>or</strong>est Service, Calif. F<strong>or</strong>. Range Exper. Stn. 17 p.<br />

MITCHELL, R.G., JOHNSON, N.E., and WRIGHT, K.H. 1974. Susceptibility of 10 spruce species and hybrids to the white<br />

pine weevil (=Sitka spruce weevil) in the Pacific N<strong>or</strong>thwest. PNW-225. U.S. Department of Agriculture, F<strong>or</strong>est<br />

Service.. 8 p.<br />

MORROW, EA., WHITHAM, T.G., POTTS, B.M., LADIGES, P., ASTON, D.H., and WILLIAMS, J.B. 1994. Gall-f<strong>or</strong>ming<br />

insects concentrate on hybrid phenotypes of Eucalyptus, p. 121-134. In Price, P.W., Mattson, W.J., and Baranchikov,<br />

Y.N., eds. The ecology and evolution of gall-f<strong>or</strong>ming insects. Gen. Tech. Rep. NC-174. U.S. Department Of Agriculture,<br />

F<strong>or</strong>est Service, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

OSAWA, A. 1989. Causality in m<strong>or</strong>tality patterns of spruce trees during a spruce budw<strong>or</strong>m outbreak. Can. J. F<strong>or</strong>. Res. 19:<br />

632-638.<br />

REMINGTON, C.L. 1968. Suture-zones of hybrid interaction between recently joined biotas, p. 321-428. In Dobzhansky,<br />

T., Hecht, M.K., and Steere, W.C., eds. Evolutionary Biology, Vol. 2. Appleton-Century-Crofts, NY.<br />

ROSE, A.H. and LINDQUIST, O.H. 1977. Insects of eastern spruces, fir and hemlock. Can. F<strong>or</strong>. Serv. F<strong>or</strong>. Tech. Rept. 23.<br />

159 p.<br />

SCHMIDT-VOGT, H. 1977. Die Fichte. Verlag Paul Parey, Berlin. 647 p.<br />

SIMMS, E.L. and FRITZ, R.S. 1990. The ecology and evolution of host-plant resistance to insects. TREE 5: 356-360.<br />

148


SINGH, D.R 1986_ Breeding f<strong>or</strong> resistance to disease and insect pests. Springer Verlag, NY. 222 p.<br />

STACE, C.A. 1987. Hybridization and the plant species, p. 115-127. In Urbanska, K.M., ed. Differentiation Patterns in<br />

Higher Plants. Academic Press, NY.<br />

TAHVANAINEN, J. and NIEMELA, R 1987. Biogeographical and evolutionary aspects of insect herbiv<strong>or</strong>y. Ann. Zool.<br />

Fenn. 24: 239-247.<br />

WHITHAM, T.G., MORROW, RA., and POTTS, B.M. 1994. Plant hybrid zones as centers of biodiversity: the herbiv<strong>or</strong>e<br />

community of two endemic Tasmanian eucalyptus. Oecologia 97: 481-490.<br />

WHITHAM, T.G. 1989. Plant hybrid zones as sinks f<strong>or</strong> pests. Science 244: 1490-1493.<br />

WRIGHT, J.W. 1955. Species crossability in spruce in relation to distribution and taxonomy. F<strong>or</strong>. Sci. 1: 319-349.<br />

YEH, EC. and ARNOTT, J.T. 1986. Electroph<strong>or</strong>etic and m<strong>or</strong>phological differentiation of Picea sitchensis, Picea glauca,<br />

and their hybrids. Can. J. F<strong>or</strong>. Res. 16:791-798.<br />

ZOBEL, B. and TALBER_I; J. t984. Applied f<strong>or</strong>est tree improvement. John Wiley and Sons, NY. 505 p.<br />

149


RECENT ADVANCES IN RESEARCH ON WHITE PINE WEEVIL ATTACNG<br />

SPRUCES IN BRITISH COLUMBIA<br />

G.K. KISS I, A.D. YANCHUK 2, R.I. ALFARO a, J.E. CARLSON 4, and J.F. MANVILLE 3<br />

1British Columbia F<strong>or</strong>est Service, Kalamalka F<strong>or</strong>estry Centre, Vernon, British Columbia, Canada<br />

2British Columbia F<strong>or</strong>est Service, <strong>Research</strong> Branch, Vict<strong>or</strong>ia, British Columbia, Canada<br />

3Canadian F<strong>or</strong>est Service, Pacific <strong>Research</strong> Centre, Vict<strong>or</strong>ia, British Columbia, Canada<br />

4University of British Columbia, Biotechnology Lab<strong>or</strong>at<strong>or</strong>y, Vancouver, British Columbia, Canada<br />

INTRODUCTION<br />

Interi<strong>or</strong> spruce (a collective term f<strong>or</strong> white spruce, Picea gIauca (Moench) Voss; Engelmann spruce, Picea<br />

engelmanni Parry, and their varying degrees of hybrids) is a very imp<strong>or</strong>tant timber resource f<strong>or</strong> the province of British<br />

Columbia. Over 100 million interi<strong>or</strong> spruce seedlings are planted annually. An estimated 1.38 billion seedlings have been<br />

planted to date, and the total number of hectares planted in spruce is over 1.3 million.<br />

White pine weevil, Pissodes strobi Peck, is a maj<strong>or</strong> cause of Sitka spruce, Picea sitchensis (Bong.) Carr., plantation<br />

failure in coastal British Columbia, Washington, and Oregon (McMullen 1976; Furniss and Carolin 1977; MacSiurtain I981,<br />

Alfaro 1982, 1989). This pest is also becoming an imp<strong>or</strong>tant menace to interi<strong>or</strong> spruce plantations, causing considerable<br />

damage to both yield and quality of product (Cozens 1983, Tayl<strong>or</strong> et al. 1991).<br />

Adult weevils overwinter in the duff layer of the f<strong>or</strong>est flo<strong>or</strong> and begin laying their eggs in the bark of young spruce<br />

trees (generally 3-30 years old) near the tip of the previous ye.ar's shoot in early summer. Young weevil larvae feed on the<br />

phloem of the leader, moving downwards in the process (Silver 1968). Once a feeding ring is f<strong>or</strong>med, the leader of the tree<br />

dies. In their downward movement, the larvae can destroy up to 4 years of growth. Literature sources refer to the successful<br />

colonization of the leader as weevil attack; <strong>this</strong> terminology will be maintained in <strong>this</strong> rep<strong>or</strong>t. Repeated leader destruction<br />

causes loss of height growth and stem def<strong>or</strong>mities, which leads to quality reduction (Alfaro 1989 and 1992).<br />

Selection of spruce varieties genetically resistant to weevil damage is a potential tool f<strong>or</strong> the reduction of weevil<br />

damage in future plantations. Genetic resistance could manifest itself in several f<strong>or</strong>ms. M<strong>or</strong>phological and anatomical<br />

differences could make certain trees less susceptible to feeding, ovipositioning, and to the development of the larvae (Plank<br />

and Gerhold 1965, Stroh and Gerhold 1965). Another possible resistance mechanism consists of variations in the chemical<br />

composition of resistant trees that make them less attractive to weevils. Phagostimulants may be lacking in some trees,<br />

which can render them less attractive to weevils. Alternately, the presence of feeding repellents, deterrents, <strong>or</strong> toxic compounds<br />

may provide an effective defense mechanism (van Buijtenen and Santamour 1972; Bridgen et al. 1979; Alfaxo 1980;<br />

Alfaro et al. 1980, 1984; Alfaro and B<strong>or</strong>den 1982, 1985; Wilkinson 1985; Brooks et al. 1987 a, b).<br />

The impact of weevil damage could be mitigated by the ability of the trees to recover from the attack (Painter 1951 ).<br />

Tolerant trees are m<strong>or</strong>e likely to overcome the effect of attacks and will suffer less growth and quality loss. This depends<br />

largely on the ability of one of the lateral branches to gain apical dominance quickly after the leader has been killed. Trees<br />

lacking <strong>this</strong> ability will develop multiple leaders (f<strong>or</strong>ks) and often are stunted.<br />

Evaluation of various trials in British Columbia provided strong evidence f<strong>or</strong> genetic variation in susceptibility f<strong>or</strong><br />

weevil damage in both Sitka spruce (Ying 1991, Alfaro and Ying 1990) and interi<strong>or</strong> spruce (Kiss and Yanchuk 1991).<br />

Mattson, W.J., NiemelL P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle.<br />

<strong>USDA</strong> F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

150


The objectives of <strong>this</strong> presentation are to (i) review measures taken by various agencies in British Columbia to<br />

elucidate the interrelationship between the weevil and its host, (ii) discuss studies in progress to take advantage of the genetic<br />

resistance of the host and (iii) present an overview of attempts to control one of the m<strong>or</strong>e destructive f<strong>or</strong>est pests. The<br />

presentation will also review future plans.<br />

Prince Ge<strong>or</strong>ge Progeny Trials<br />

WEEVIL RESEARCH<br />

Field Studies<br />

Due to the imp<strong>or</strong>tance of interi<strong>or</strong> spruce in British Columbia, a decision was made to embark on the genetic improvement<br />

of the species complex in the early 1960's. The seni<strong>or</strong> auth<strong>or</strong> was hired to carry out <strong>this</strong> project in 1967.<br />

The basic approach to improving interi<strong>or</strong> spruce was to designate geographically distinct areas (selection units <strong>or</strong><br />

SU's) and select trees based on size and f<strong>or</strong>m in each selection unit (Kiss and Yeh 1988, Kiss and Yanchuk 1991). The first<br />

of these selection units was in the Prince Ge<strong>or</strong>ge area. A total of 180 parent trees were selected. Cones and scions were<br />

collected from most selected parents. Scions were used to establish clone banks, and seeds were used to establish genetic<br />

tests.<br />

As part of the improvement program, genetic tests were established using open-pollinated progenies of individual<br />

parents selected in various selection units. The first set of these tests, representing the Prince Ge<strong>or</strong>ge SU, was established in<br />

1972 and 1973 (Kiss and Yeh 1988, Kiss and Yanchuk 1991) using the open-pollinated progenies of 173 parent trees.<br />

The initial objective of the genetic tests was to identify well-perf<strong>or</strong>ming families based on the growth perf<strong>or</strong>mance of<br />

their progenies. This inf<strong>or</strong>mation aided in the construction of rogued first generation seed <strong>or</strong>chards and helped to identify<br />

candidates f<strong>or</strong> advanced generation breeding. To <strong>this</strong> end, periodic height measurements were carried out and at age 15<br />

diameter measurements were also made.<br />

Following the 15-year height and diameter assessments, heavy infestation of white pine weevil was observed at some<br />

test sites. Curs<strong>or</strong>y observation appeared to suggest a pattern to the infestation. Further studies confirmed <strong>this</strong> pattern and<br />

revealed the possibility of genetically controlled resistance mechanism(s) in a number of families. Results of <strong>this</strong> study are<br />

rep<strong>or</strong>ted by Kiss and Yanchuk (1991).<br />

In summary, the results showed that:<br />

1. While there were different intensities of attack at the three individual test sites in the Prince Ge<strong>or</strong>ge SU tests<br />

(i.e., 9% at Quesnel, 37% at Red Rock, and 63% at Aleza Lake), there were significant c<strong>or</strong>relations between<br />

families attacked across sites (Figs. 1 and 2).<br />

2. The estimate of family-mean heritability f<strong>or</strong> weevil resistance across sites was high (h2f=0.77 + 0.11).<br />

3. Families exhibiting greater vig<strong>or</strong> at any age were generally m<strong>or</strong>e resistant. C<strong>or</strong>relation between 6-year family<br />

height and incidence of weevil attack was negative and significant (Fig. 3, r=-0.42), as well as 15-year height<br />

and incidence of weevil attack (Fig. 4, r=-0.68). In fact, when families were ranked based on height growth at<br />

different ages, and average attack was calculated f<strong>or</strong> the top and bottom 25% of families, the top families<br />

always had significantly lower attacks than those of the bottom families (Table 1).<br />

151


30 1 O0 . ._.--,,=<br />

A _ A<br />

r = 0.63 • A O_ r = 0.71 • • • • • • 2 •<br />

. m 80. - A _• _ _ • •<br />

3 • • N A<br />

• _iA'_A%A "<br />

• A A • A < _ 60. A• A_ AA} _AA &A •i A•A• A ••<br />

• A m A A •_ t<br />

"_ 15 AA A A _A • --_ A_ _A• A •<br />

© • • A A • • O_ • A AA•• _ A •<br />

>" 10 AA • • _ • AAA • A<br />

E<br />

A• A A AA_ _MkaAA A• -- A<br />

=<br />

A A •_ • A _i_ • ,4.._ 204 _ •<br />

c 5 •,_ ••_ AaA_ • =Aj<br />

• _A SA• AA A_A •A • •<br />

o o l<br />

0 20 40 60 80 0 20 40 60 80<br />

Mean family attack at Red Rock (%) Mean family attack at Red Rock {%)<br />

Figures 1 & 2.--Comparison of attack percentages at the Red Rock and Aleza Lake test sites and at the Red Rock and<br />

Quesnel test sites.<br />

8O<br />

• 80,.<br />

0 • A • •<br />

60 A A 2 •A d O0 a A •<br />

c:?-. • " A.<br />

'--- • _ • •_•_ #*A • _a > • •a_a AA • •<br />

• _A A 0.68 A a • •_2 A a<br />

. • _ r=-<br />

0 , r_ , •<br />

' ' ' 0 _ .......... A A A<br />

50 55 60 65 70 75 80 85 90 160 180 200 220 240 260<br />

Mean 6-year heights at Red Rock (%) o 300 32<br />

Mean 15-year heights at Red Rock (%)<br />

Figures 3 & 4.--Relationship between average family height measurements at 6 and 15 years of age and mean family weevil<br />

attacks at the Red Rock test site.<br />

This latter finding was especially gratifying as a literature survey indicated that white pine weevil fav<strong>or</strong>ed m<strong>or</strong>e<br />

robust, longer leaders. While <strong>this</strong> still may be true at the phenotypic level, genetically superi<strong>or</strong> families are less susceptible.<br />

We would have accepted susceptibility that was equal to the mean; the fact that faster growing families are less susceptible is<br />

a bonus.<br />

Quesnel Lakes Progeny Trials<br />

In the fall of t993 additional genetic test plantations were assessed f<strong>or</strong> weevil attack. These plantations were<br />

established in 1983 to evaluate the genetic potential of I42 parent trees. The tests were established on three sites (KetchuE_<br />

Creek, Little Benson Lake, and Clearwater) using randomized complete blocks with eight blocks per site and four seedlings<br />

per replication per family (total 32 seedlings per site per family f<strong>or</strong> a grand total of 96 seedlings per family over the three<br />

sites).<br />

152


Table 1.--Average weevil attack on the top and bottom 25% of families at three test sites. Families were ranked by average<br />

heights at various ages and at 15-yr dbh (diameter at breast height).<br />

Test age<br />

Quesnel test Red Rock test Aleza Lake test<br />

Top Bot. Diff. Top Bot. Diff. Top Bot. Diff.<br />

Initial ht. 6.6 10.7 4.1 31.0 42.9 11.9 56.6 68.3 11.7<br />

3-year ht. 7.1 11.9 4.8 30.6 44.4 13.8 55.9 69.4 13.5<br />

6-year ht. 6.6 12.1 5.5 28.8 45.1 16.3 53.6 71.3 17.7<br />

10-year ht. 5.9 11.9 6.0 26.4 45.0 18.6 52.7 70.8 18.1<br />

15-year ht. 5.2 13.3 8.1 23.8 49.0 25.2 50.6 73.3 22.7<br />

DBH 6.3 11.7 5.4 28.3 44.1 15.8 55.6 70.1 14.5<br />

Moderate to heavy weevil infestation occurred on two of these plantations (Little Benson Lake - 17%, Clearwater -<br />

33%). As was the case with the earlier studies, the weevils again fav<strong>or</strong>ed certain families and avoided others. Some individuals<br />

were attacked up to five times over the years, while m<strong>or</strong>e than 1,700 of the 2,700 intensively studied trees at<br />

Clearwater have never been attacked. There was high c<strong>or</strong>relation between mean family attacks at the two sites (Fig. 5, r<br />

=0.66).<br />

A<br />

70 A<br />

60 A<br />

50 AA _AA •<br />

-_ O • AAI•<br />

AA • •<br />

23o _ • I i. ••_<br />

_ 20 _, A• • • •<br />

r = 0.66<br />

10 • AA<br />

0 10 20 30 40 50 60<br />

Mean family attack at Little Benson(%)<br />

Figure 5.--Comparison of attack percentages at Little Benson and Clearwater.<br />

Once again the m<strong>or</strong>e vig<strong>or</strong>ous, faster growing families were less susceptible to weevil attacks. Based on the overall<br />

ranking at all sites, the top 25% of families suffered substantially less damage than the bottom 25% (Table 2).<br />

Additional data collected on the Clearwater site in cooperation with Dr. Ren6 Alfaro relate to number of attacks per<br />

tree, recovery of attacked trees, and quality traits following recovery. These data are presently being evaluated and the<br />

results will be published elsewhere.<br />

A<br />

153


Table 2.--Average weevil attack on the top and bottom 25% of families at the Little Benson and Clearwater test sites.<br />

Families were ranked by average heights at various ages.<br />

Test age<br />

Comparative Trials<br />

Little Benson test Clearwater test<br />

Top Bot. Diff. Top Bot. Diff.<br />

Initial ht. 15.5 18.2 2.7 28.7 40.8 12.1<br />

3-year ht. 14.4 18.6 4.2 26.3 41.4 15.1<br />

6-year ht. 10.4 19.7 9.3 22.4 43.8 21.4<br />

10-year ht. 7.6 21.8 14.2 17.5 47.6 30.1<br />

In 1992 another pilot project plantation was evaluated f<strong>or</strong> weevil damage. This plantation was established in 1976 to<br />

compare the relative perf<strong>or</strong>mance of native British Columbian spruces with those <strong>or</strong>iginating from eastern Canada. Eastern<br />

spruces up to now perf<strong>or</strong>m well in B.C., and we are planning to inc<strong>or</strong>p<strong>or</strong>ate some eastern material into our breeding program.<br />

This plantation is a replicated complete block design with 15 blocks, 10-tree row plots in each block from each of 21<br />

families. There are 9 families from eastern Canada (ENA) and 12 families from B.C. (four each from East Kootenay (EK),<br />

Prince Ge<strong>or</strong>ge (PG), and Prince Rupert (PR)).<br />

Results indicate that eastern white spruce is very resistant to the white pine weevil. Of the 1,220 eastern trees, only<br />

67 were damaged by weevil, whereas 379 native spruces were damaged out of a total of 1,665 (Table 3). Interestingly, 25 of<br />

the 67 individuals damaged were progenies of the same parent. These results are very significant in light of the fact that these<br />

trees also grow and survive extremely well.<br />

Table 3.--Weevil attack differences among families of different geographic <strong>or</strong>igin.<br />

Parents Number of trees<br />

Source Number Alive Attacked Percent<br />

ENA 9 1,220 67 5.5<br />

EK 4 545 154 28.3<br />

PG 4 562 85 15.1<br />

PR 4 558 140 25.1<br />

BC Total 12 1,665 379 22.8<br />

Further proof of the resistance is demonstrated by the eastern N<strong>or</strong>th American white spruce clones (about 75 clones)<br />

located in our breeding arb<strong>or</strong>eta at the Kalamalka F<strong>or</strong>estry Centre. These clones are surrounded by native B.C. clones that<br />

are heavily attacked by weevil yet eastern clones are almost void of attacks. It is not possible to treat these data using<br />

statistical procedures due to the small number of ramets per clone (4), but there is no doubt about the trend.<br />

DISCUSSION<br />

Based on our observations, we must review weevil behavi<strong>or</strong> on interi<strong>or</strong> spruce. It has been postulated that in Sitka<br />

spruce, white pine weevil preferentially attacked the most vig<strong>or</strong>ous, longest, and thickest leaders (VanderSar and B<strong>or</strong>den<br />

154


1977, Kline and Mitchell 1979, Wood and McMullen 1983). It was suggested that the same preference holds f<strong>or</strong> interi<strong>or</strong><br />

spruce as well. Our investigation confirms beyond doubt that it is not the case. Susceptibility of the individual is the<br />

overriding principle that determines whether <strong>or</strong> not it will be attacked successfully.<br />

In field studies we observed that some seedlings of susceptible families, measuring 50 cm, were attacked by weevil.<br />

In some cases these young seedlings were attacked below the previous year's growth probably due to the inadequate thickness<br />

of the previous year's stem.<br />

To clarify weevil behavi<strong>or</strong>, a number of studies are being initiated to elucidate the biology of white pine weevil in the<br />

interi<strong>or</strong> and its relationship to its host. Verification of resistance in putatively resistant families will also be carried out.<br />

During the spring of 1993 a number of crosses were made using various putatively weevil resistant and susceptible<br />

parents. Judgement as to resistancy <strong>or</strong> susceptibility was made based on the progenies' weevil resistance perf<strong>or</strong>mance.<br />

These crosses will f<strong>or</strong>m the basis of a number of experiments designed to verify resistancy and susceptibility and to study the<br />

mode and degree of inheritance of these traits.<br />

Future plans include continuation of the invent<strong>or</strong>y of resistant families, provision of material f<strong>or</strong> other agencies, and<br />

inc<strong>or</strong>p<strong>or</strong>ation of the accumulated knowledge into the breeding program.<br />

Terpene Studies<br />

Lab<strong>or</strong>at<strong>or</strong>y Studies<br />

Our first intention following the identification of genetically controlled resistance was to try to identify the reason f<strong>or</strong><br />

resistance. To <strong>this</strong> aim, we enlisted the help of a team of biochemists lead by Dr. John Manville. They tested eight resistant<br />

and eight susceptible parental clones f<strong>or</strong> terpenes that they judged to be of potential use in differentiating between them. The<br />

tests included both leaf and bark samples.<br />

They were able to classify 15 of the 16 parental clones as to susceptibility <strong>or</strong> resistance using six bark terpenes with<br />

multivariate technique (Manville et al. 1994). Two of the terpenes are identified (santene and citronellyl acetate) the other<br />

four are not yet identified.<br />

Leaf terpenes also provided similar results. Using six leaf terpenes (terpinolene and five unidentified terpenes), they<br />

were able to classify c<strong>or</strong>rectly 15 of 16 samples.<br />

Future plans include validation of the results (i.e., using a random sample of attacked and unattacked trees and<br />

applying the technique to them). They are also w<strong>or</strong>king on the identity of the unidentified terpenes.<br />

DNA Marker Studies<br />

We were also interested in investigating the potential differences at the gene level. To study these differences, we<br />

enlisted the help of Dr. John E. Carlson of the University of British Columbia (Carlson et al. 1994).<br />

The team investigated differences between resistant and susceptible interi<strong>or</strong> spruce parents and their open-pollinated<br />

progenies. They constructed genetic linkage maps based on the recently developed Random Amplified Polym<strong>or</strong>phic DNA<br />

(RAPD) marker system, the first linkage maps available f<strong>or</strong> any spruce species (Hong et al. 1993). Using a "Bulk Parental<br />

Analysis" (BPA) technique developed by the team, they pooled DNA from 12 resistant and 12 susceptible parents. This<br />

technique identified 20 DNA markers putatively associated with weevil resistance.<br />

The followup investigation was carried out on half-sib progenies of the <strong>or</strong>iginal parents. Needles were collected from<br />

10 half-sib individuals of each parent of the previous study (a total of 240 half-sib progenies).<br />

155


Results of these studies showed that three of the markers were strongly c<strong>or</strong>related with resistance. None of the<br />

markers showed 100% association with resistance <strong>or</strong> susceptibility. The seni<strong>or</strong> auth<strong>or</strong> has two potential explanations f<strong>or</strong> <strong>this</strong>:<br />

1. The markers are only loosely linked to the gene(s) f<strong>or</strong> resistance, so recombination between the marker and the<br />

gene can still take place.<br />

2. M<strong>or</strong>e than one gene may be involved in weevil resistance, and/<strong>or</strong> modifier genes are also necessary f<strong>or</strong> a<br />

resistance gene to be expressed.<br />

These studies are also continuing.<br />

Micropropagation of Resistant Genotypes<br />

Dr. Ben Sutton and his team from the British Columbia F<strong>or</strong>est Biotechnology Centre in Vancouver are involved in<br />

utilizing elite full-sib crosses among resistant parents by multiplying them through embryogenesis (Roberts 1994). Their<br />

technique to produce large numbers of"emblings" has been developed to a point where they can mass produce these<br />

propagules.<br />

Embryogenic tissues can be st<strong>or</strong>ed in liquid nitrogen indefinitely. Once clones have been tested and proven to be<br />

resistant, they can be mass produced and used f<strong>or</strong> operational plantations.<br />

Results of the studies described above will be utilized in developing an integrated weevil control system that will<br />

provide protection to newly planted spruce f<strong>or</strong>ests in British Columbia. There are a number of unanswered questions that<br />

will be addressed in future studies, but the results to date are encouraging. We are optimistic that our eff<strong>or</strong>ts will lead to<br />

success.<br />

SUMMARY<br />

Studies of interi<strong>or</strong> spruce progeny trials revealed apparent genetic control of resistance to weevil damage in interi<strong>or</strong><br />

spruce in British Columbia. C<strong>or</strong>relation between mean family attacks across sites was high.<br />

Contrary to previous suggestions, weevil damage was m<strong>or</strong>e prevalent on less vig<strong>or</strong>ous families. The difference<br />

between average damage of higher and lower ranked families (based on growth perf<strong>or</strong>mance at any age) was always significant.<br />

Families that ranked highest f<strong>or</strong> growth typically suffered less damage.<br />

Eastern N<strong>or</strong>th American white spruce appears significantly m<strong>or</strong>e resistant to weevil than are western sources.<br />

Western sources suffered m<strong>or</strong>e than fourfold the damage of eastern sources.<br />

Lab<strong>or</strong>at<strong>or</strong>y studies appeared to confirm differences between putatively resistant and susceptible spruces:<br />

I. Terpene analyses tentatively identified terpene content differences between the two strains. Fifteen of sixteen<br />

families could be distinguished by <strong>this</strong> technique;<br />

2. RAPD markers could also differentiate between resistant and susceptible families. While none of the markers<br />

had 100% association with resistance, the results are clearly pointing to markers loosely associated with<br />

resistance.<br />

Genotypes identified as resistant will be propagated using a technique referred to as somatic embryogenesis. A team<br />

at the British Columbia <strong>Research</strong> Inc<strong>or</strong>p<strong>or</strong>ation has perfected the technique so that it can mass produce clones that are<br />

resistant to weevil.<br />

156


Further research is underway to clarify the many questions that are still remaining. Nevertheless, the resistance<br />

observed in our genetic trials will allow us to select and breed f<strong>or</strong> <strong>this</strong> attribute in combination with growth and other<br />

desirable traits.<br />

ACKNOWLEDGEMENTS<br />

The auth<strong>or</strong>s are indebted to many individuals who contributed to <strong>this</strong> manuscript. Technical supp<strong>or</strong>t was provided by<br />

D. Wallden and G. Phillips over many years. The financial supp<strong>or</strong>t of the British Columbia F<strong>or</strong>est Service was responsible<br />

f<strong>or</strong> the establishment, maintenance, and evaluation of the trials that provided the database f<strong>or</strong> the study.<br />

LITERATURE CITED<br />

ALFARO, R.I. 1980. Host selection by Pissodes strobi Peck: Chemical interaction with the host plant. Ph.D. Thesis, Simon<br />

Fraser University, Vancouver, British Columbia.<br />

ALFARO, R.I. 1982. Fifty-year-old Sitka spruce plantations with a hist<strong>or</strong>y of intensive weevil attack. J. Entomol. Soc.<br />

British Columbia 79: 62-65.<br />

ALFARO, R.I. 1989. Stem defects in Sitka spruce induced by Sitka spruce weevil, Pissodes strobi Peck, p. 177-185. In:<br />

Alfaro, R.I., and S. Glover, eds. Insects Affecting Ref<strong>or</strong>estation: Biology and Damage. Proceedings of a IUFRO<br />

symposium held on July 3-9, 1988, in Vancouver, B.C., Canada, under the auspices of the XVIII International<br />

Congress of Entomology. F<strong>or</strong>estry Canada, Vict<strong>or</strong>ia, B.C.<br />

ALFARO, R.I., PIERCE H.D., Jr., BORDEN, J.H., and OEHLSCHLAGER, A.E. 1980. Role of volatile and nonvolatile<br />

components of Sitka spruce bark as feeding stimulants f<strong>or</strong> Pissodes strobi Peck (Coleoptera: Curculionidae). Can J.<br />

Zool. 58: 626-632.<br />

ALFARO, R.I., BORDEN, J.H., HARRIS, L.J., NIJHOLT, W.W., and MCMULLEN, L.H. 1984. Pine oil, a feeding<br />

deterrent f<strong>or</strong> the white pine weevil Pissodes strobi Peck (Coleoptera: Curculionidae). Can. Ent. 116: 41-44.<br />

ALFARO, R.I. and B<strong>or</strong>den, J.H. 1982. Host selection by the white pine weevil, Pissodes strobi Peck: feeding bioassays<br />

using host and nonhost plants. Can. J. F<strong>or</strong>. Res. 12: 64-70.<br />

ALFARO, R.I. and BORDEN, J.H. 1985. Fact<strong>or</strong>s determining the feeding of the white pine weevil on its coastal British<br />

Columbia host, Sitka spruce. Proc. Ent. Soc. Ont. 116 (supplement): 63-66.<br />

ALFARO, R.I. and YING, C.C. 1990. Levels of Sitka spruce weevil, Pissodes strobi Peck, damage among Sitka spruce<br />

provenances and families near Sayward, British Columbia. Can. Ent. 122: 607-615.<br />

BRIDGEN, M.R., HANOVER, J.W., and WILKINSON, R.C. 1979. Ole<strong>or</strong>esin characteristics of eastern white pine seed<br />

sources and relationship to weevil resistance. F<strong>or</strong>. Sci. 25: 175-183.<br />

BROOKS, J.E., BORDEN, J.H., PIERCE, H.D., JR., and LISTER, G.R. 1987a. Seasonal variation in foliar and bud<br />

monoterpenes in Sitka spruce. Can. J. Bot. 65: 1249-1252.<br />

BROOKS, J.E., BORDEN, J.H., and Pierce, H.D., Jr. 1987b. Foliar and c<strong>or</strong>tical monoterpenes in Sitka spruce: potential<br />

indicat<strong>or</strong>s of resistance to the white pine weevil, Pissodes strobi Peck (Coleoptera: Curculionidae). Can J. F<strong>or</strong>. Res.<br />

17: 740-745.<br />

CARLSON, J.E., HONG, YONG-PYO, and KISS, G. 1994. DNA markers associated with weevil resistance in interi<strong>or</strong><br />

spruce. Presented at the National white pine weevil w<strong>or</strong>kshop Richmond, British Columbia, January 19-21. In press.<br />

157


COZLN_, R.D. 1983. The spruce weevil, Pissodes strobi Peck (Coleoptera: Curculionidae). A review of its biology,<br />

damage and control techniques with reference to the Prince Ge<strong>or</strong>ge Timber Supply area. B.C. Ministry of F<strong>or</strong>ests<br />

Prince Ge<strong>or</strong>ge. F<strong>or</strong>. Serv. Intern. Rep. PM-PG-3.<br />

I::URNISS, R.L. and CAROLIN, V.M. 1977. Western f<strong>or</strong>est insects. Misc. Pu. 1339. Washington, DC: U.S. Department of<br />

Agriculture, F<strong>or</strong>est Service.<br />

_IONG, YONG-PYO, GLAUBITZ J.C., PONOY, B., and CARLSON, J.E. 1992. DNA finger<strong>print</strong>ing of conifer genomes<br />

using RAPD markers, p. 8-22. In Stomp, A.M., ed. Proc. Intl. Conifer Biotech W<strong>or</strong>king Group Meeting, April<br />

23-28, 1992, Raleigh, N<strong>or</strong>th Carolina.<br />

KISS, G.K. and YANCHUK, A.D. 1991. Preliminary evaluation of genetic variation of weevil resistance in interi<strong>or</strong> spruce<br />

in British Columbia. Can. J. F<strong>or</strong>. Res. 21: 230-234.<br />

KISS_ G. and YEH, F.C. 1988. Heritability estimates f<strong>or</strong> height f<strong>or</strong> young interi<strong>or</strong> spruce in British Columbia. Can. J. F<strong>or</strong>.<br />

Res_ 18: 158-162.<br />

KHNE, L.N. and MITCHELL, R.G. 1979. Insects affecting twigs, terminals and buds. In Rudinsky, J.A., ed. F<strong>or</strong>est Insect<br />

Survey and Control. O.S.U. Book St<strong>or</strong>es, Inc.<br />

MACSIURTAIN, M.P. 1981. Distribution, management, variability and economics of Sitka spruce (Picea sitchensis<br />

(Bong.) Cart.) in coastal British Columbia. M.Sc thesis. University of B.C., Faculty of F<strong>or</strong>estry, Vancouver, B.C.<br />

MANVILLE, J.F., NAULT, J., VON RUDLOFF E., YANCHUK, A., and KISS, G.K. 1994. Spruce terpenes: Expression<br />

and weevil resistance. Presented at the National white pine weevil w<strong>or</strong>kshop Richmond, British Columbia, January<br />

19-21. In press.<br />

PAINTER, R.H. 1951. Insect resistance in crop plants. MacMillan Pub. Co. New Y<strong>or</strong>k. 520 p.<br />

PLANK, G. H. and GERHOLD, H.D. 1965. Evaluating host resistance to the white pine weevil, Pissodes strobi, using<br />

feeding preference tests. Ann. Ent. Soc. Am. 58: 527-532.<br />

ROBERTS, D.R. 1994. Somatic embryogenesis f<strong>or</strong> mass propagation of weevil resistant spruce. Presented at the National<br />

white pine weevil w<strong>or</strong>kshop Richmond, British Columbia, January 19-21. In press.<br />

SILVER, G.T. 1968. Studies on the Sitka spruce weevil, Pissodes sitchensis, in British Columbia. Can. Ent. 100: 93-110.<br />

STROH, R.C. and GERHOLD, H.D. 1965. Eastern white pine characteristics related to weevil feeding. Silvae Genetica 14:<br />

160-169.<br />

TAYLOR, S., ALFARO, R.I., and LEWIS, K. 1991. Fact<strong>or</strong>s affecting the incidence of white pine weevil damage to white<br />

spruce in the Prince Ge<strong>or</strong>ge Region of British Columbia. J. Entomol. Soc. B.C. 88: 3-7.<br />

VAN BUIJTENEN, J.P. and SANTAMOUR, F.S. 1972. Resin crystallization related to weevil resistance in white pine<br />

(Pinus strobus). Can. Ent. 104: 215-219.<br />

VANDERSAR, T.J.D. and BORDEN, J. H. 1977. Visual <strong>or</strong>ientation of Pissodes strobi Peck (Coleoptera: Curculionidae) in<br />

relation to host selection behaviour. Can. J. Zool. 55: 2042-2049.<br />

WILKINSON, R.C. 1985. Comparative white-pine weevil attack susceptibility and c<strong>or</strong>tical monoterpene composition of<br />

western and eastern white pines. F<strong>or</strong>. Sci. 31: 39-42.<br />

WOOD, R.O. and MCMULLEN, L.H. 1983. Spruce weevil in British Columbia. Can. F<strong>or</strong>. Serv. Pac. F<strong>or</strong>. Res. Cent. F<strong>or</strong>.<br />

Pest Leafl. 2.<br />

YING, C.C. 1991. Genetic resistance to the white pine weevil in Sitka spruce. B.C. Min. of F<strong>or</strong>. Res. Note 106. Vict<strong>or</strong>ia,<br />

B_C.,Canada. 17 p.<br />

158


A SUGI CLONE HIGHLY PALATABLE TO HARES:<br />

SCREENING FOR BIOAC_IIVE CHEMICA S<br />

HIROFUMI ttlRAKAWA, TADAKAZU NAKAStIIMA, YOSttlOKI HAYASHI,<br />

SEIJI OHARA, and TSUTOMU KUWAHATA<br />

F<strong>or</strong>estry and F<strong>or</strong>est Products Resealch Institute<br />

PO Box 16, Tsukuba-N<strong>or</strong>in, Ibaraki 305, Japan<br />

INTRODUCTION<br />

Herbiv<strong>or</strong>es often selectively feed on some plant individuals over other conspecifics. Such differences in palatability<br />

are usually chemically determined. Although typically confounded with environmental fact<strong>or</strong>s (Rousi et al. 1989, 1991,<br />

1993), genetic control of the differences in palatability observed among geographic <strong>or</strong>igins (Matsuura and Maeda 1981;<br />

Radwan et al. 1982; Hansson et al. 1986; Rousi et al. 1989, 1991), among families (Rousi et al. 1989, 1991), and among<br />

clones (Radwan 1972, Hara 1973, Dimock et al. 1976, Kuwahata and Tanbara 1983) is obvious.<br />

Although mammalian herbiv<strong>or</strong>es can cause severe damage to trees, it is usually difficult to control such animals,<br />

because of economical, aesthetic, and ethical reasons. Breeding f<strong>or</strong> resistance is one of the possible solutions to <strong>this</strong> problem.<br />

Theref<strong>or</strong>e, eff<strong>or</strong>ts have been made to find the chemical agents responsible f<strong>or</strong> the variation in palatability among trees.<br />

Atetsu 4 (A4), an elite sugi clone, Cryptomeria japonica, which <strong>or</strong>iginated in Atetsu, Okayama, Japan, is extremely<br />

palatable to hares. Its high palatability was first observed at a test planting site of 32 sugi clones, where A4 was highly<br />

damaged by hares just 1 year after planting. Later, an unusually high m<strong>or</strong>tality rate was found in A4 at another test site of 37<br />

clones 5 years after planting, and damage by hares were considered to be the main cause of the m<strong>or</strong>tality (Kuwahata and<br />

Tanbara 1983).<br />

This paper reconfirms the high palatability of A4 to hares, analyzes some of the chemical properties of the three sugi<br />

clones, and tests the palatability of crude chemical extracts and separated fractions using filter paper trials.<br />

MATERIALS AND METHODS<br />

Plant Materials and Hares<br />

The palatable sugi clone, A4, and two other unpalatable sugi clones, Uda 36 (U36) and Niimi 4 (N4), were used. The<br />

clones were transplanted to the nursery of the F<strong>or</strong>estry and F<strong>or</strong>est Products <strong>Research</strong> Institute in Ibaraki, and leaves were<br />

collected from trees at each feeding trial and, when stocked f<strong>or</strong> use in chemical analysis, frozen at -43°C. Captive hares,<br />

Lepus brachyurus, kept in pens at the institute, were used f<strong>or</strong> palatability tests. The hares were outdo<strong>or</strong>s and maintained on<br />

commercial food pellets. They were either b<strong>or</strong>n in captivity <strong>or</strong> captured in the wild soon after birth.<br />

Palatability Test<br />

Feeding trials were conducted using two hares in separate 3.6 x 3.6 m pens. Fresh leaves <strong>or</strong> extracted residues of the<br />

clones were put on shallow mesh trays (24 x 32 x 1cm) and offered to the hares. Consumption was monit<strong>or</strong>ed by measuring<br />

the weight of the remnants over several days. C<strong>or</strong>rections were made f<strong>or</strong> changes in leaf weight due to dehydration by<br />

measuring controls.<br />

Mattson, W.J., Niemel_i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle.<br />

<strong>USDA</strong> F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

159


Filter paper tests were used to examine the palatability of fractions. Of six pieces of filter paper (5 x 8 cm, 1 g), two<br />

were treated with the extract of A4 and two with that of N4; the remaining two were left untreated as a control. The amount<br />

of fractions applied to each piece of filter paper was a one-gram dry-leaf equivalent. These were offered to hares kept in<br />

cages (80 x 80 x 80 cm) <strong>or</strong> pens (1.8 x 3.6 m), f<strong>or</strong> one night and the amount (area) of filter paper left uneaten was measured.<br />

Commercial food pellets were always available during the trials, hence the intake of leaves <strong>or</strong> filter paper by hares<br />

was entirely voluntary.<br />

Seasonal Stability of Palatability and Solvent Extracts<br />

Seasonal stability of the palatability of fresh leaves of A4, U36, and N4 was investigated by feeding trials every 3<br />

months from May 1988 to May 1989. Leaves were taken from the trees and cut into approximately 5 to 10cm lengths. The<br />

leaves of each clone were put on two trays, which were set at opposite positions in a pen to prevent biased consumption due<br />

to tray position.<br />

F<strong>or</strong> the same materials, solvent extracts were investigated: essential oils were taken by steam distillation f<strong>or</strong> 10 h;<br />

dichl<strong>or</strong>omethane and methanol extracts were consecutively obtained using a Soxhlet extract<strong>or</strong>. Leaves were cut into small<br />

pieces (< 3 mm in length) with pruning shears bef<strong>or</strong>e the analysis.<br />

Effects of Extraction on Palatability<br />

Two series of consecutive extractions were made: (a) steam distillation f<strong>or</strong> I0 h followed by methanol extraction; and<br />

(b) extraction by n-hexane and then methanol. The latter was done in a Soxhlet extract<strong>or</strong>. Leaves were cut into small pieces<br />

(ca. 3 mm in length) with pruning shears bef<strong>or</strong>e extraction.<br />

Hot-water extraction was made using a common kitchen pot (capacity: 5 1) and a large steam-heated stainless pot f<strong>or</strong><br />

mass extraction (capacity: 50 1). Leaves of 10 to 20 cm in length were used. Leaves in the kitchen pot were boiled f<strong>or</strong> 7 days<br />

(6 h per day) with water changed twice a day. In the steam-heated pot, leaves were boiled f<strong>or</strong> 8, 16, <strong>or</strong> 24 h and once f<strong>or</strong> 5<br />

days with water changed daily.<br />

The palatability of residual leaves were investigated by cafeteria test. Comparisons of palatability were made among<br />

the clones after each extraction.<br />

Effects of Methanol Extracts on Fresh and Residual Leaves<br />

The fresh leaves and residue after extraction of A4 and N4 were treated with the methanol extracts of A4 and N4 and<br />

feeding trials were made to compare the palatability bef<strong>or</strong>e and after treatment. Methanol extracts were taken by soaking<br />

dried fresh leaves (600 g wet weight) first in 36 1 methanol f<strong>or</strong> 13 days and then 18 1methanol f<strong>or</strong> 8 days. The extracts were<br />

combined and concentrated in a vacuum evap<strong>or</strong>at<strong>or</strong> to a small volume. The concentrate was then sprayed on fresh leaves <strong>or</strong><br />

leaf residue with the amount that was <strong>or</strong>iginally included in the fresh leaves.<br />

Palatability of Extracts<br />

Fresh leaves of 10 to 20 cm in length (260 g wet weight) were put in large flask (capacity: 10 1)with methanol and<br />

heated in a hot water bath. The 2 h extraction was repeated twice f<strong>or</strong> the same material each time with 7 1 methanol. The<br />

extracts were then combined (MW).<br />

The combined extract (MW) was added to 500 ml water and then evap<strong>or</strong>ated to approximately 400 ml, and the waterinsoluble<br />

material was precipitated. Water-soluble (W) and -insoluble (M) fractions were separated by decanting. The water'<br />

soluble fraction was then consecutively extracted with n-hexane (Wh), ethyl acetate (We), and butanol (Wb) to leave a<br />

residual water fraction (Ww). The water-insoluble fraction was dissolved with methanol and then extracted with n-hexane<br />

(Mh) to leave a residual methanol fraction (Mm).<br />

The palatability of the various fractions was examined by filter paper tests. Each fraction was tested in pairs of A4<br />

and N4 ....<br />

160


RESULTS<br />

Seasonal Stability of Palatability and Solvent Extracts<br />

A4 showed highpalatabilitythroughout theyear andwas completelyconsumedin thefirst two nights with only one<br />

exception (Fig. 1). In contrast, U36 andN4 were not eatenmucheven afterA4 was depleted. In one instance,however, U36<br />

waseaten afterA4 was depleted although N4 did not show any such tendency. The two hares showed a notablesimilarityin<br />

their feeding.<br />

g 6O<br />

4O<br />

2<br />

Atetsu 4 Uda 36 Niimi 4<br />

May 1988 Aug. 1988 Nov. 1988 Feb. 1989 May 1989<br />

, J | i i<br />

g 60 "[...<br />

40<br />

20<br />

O • o-_¢--_o_<br />

0 1 23 4 01 23 4 01 234 0 1 23 4 0 1 23 4<br />

Days of exposure<br />

Figure 1.--Palatability of fresh leaves of three sugi clones indifferent seasons. Vertical axis represents the weight of<br />

uneaten leaves remaining. Upper and lower graphs each represent the results f<strong>or</strong> one hare.<br />

The mass of essential oils in A4 was much less than that in U36<strong>or</strong> N4 throughout the year (Fig. 2). Also, the amount<br />

of dichl<strong>or</strong>omethane extract from A4 was much less than that from the other two unpalatable clones in all seasons. In contrast,<br />

the amount of methanol extract was not much different among the three clones.<br />

Atetsu 4 Uda 36 Niimi 4<br />

O O-- --,--<br />

a: Essential oils b: Dichl<strong>or</strong><strong>or</strong>nethane c: Methanol extract (%)<br />

32 /__(ml_lOOgd'm') 108 306_ 4z<br />

extract (%) 40 20 _*<br />

0 May Aug.Nov. Feb. May May Aug.Nov. Feb. May May Aug.Nov. Feb. May<br />

Figure 2.--Seasonal fluctuation of essential oils and solvent extracts in three sugi clones.<br />

161


Effects of Extraction on Palatability<br />

When boiled in a common kitchen pot f<strong>or</strong> 7 days, the residual leaves of A4 still retained high palatability. However,<br />

boiled in the steam-heated pot f<strong>or</strong> 5 days, A4 lost its palatability to hares, although it retained high palatability after boiling<br />

f<strong>or</strong> up to 24 h. The percent dry matter of the fresh leaves used in these trials was 35-39%; after boiling, residual leaves were<br />

22-25% of the <strong>or</strong>iginal fresh weight. Thus, about one-third of the dry matter was extracted by boiling. However, when<br />

boiled in the steam-heated pot f<strong>or</strong> 5 days, the residue of A4 was reduced in weight to 18%of the <strong>or</strong>iginal, whereas N4<br />

remained at 24%.<br />

When essential oils were removed by steam distillation, A4 retained its palatability, but further extraction with<br />

methanol made it unpalatable (Fig. 3). Similarly, after extraction with n-hexane, A4 residue did not lose palatability but<br />

further extraction with methanol rendered it unpalatable. In contrast, the unpalatability of U36 and N4 seemed to diminish<br />

after each extraction.<br />

Atetsu 4 Uda 36 Niimi 4<br />

a: Fresh leaves b:Steam-distilled +MeOH-extracted c:Hexane-extracted+MeOH-extracted<br />

0 t , , .. , ,<br />

_ _ I I 1 I _'I_.l_l, .ll[1._ * __l.J<br />

O I Z 3 0 I Z 3 0 I Z 3 0 I Z 3 0 I 2 3<br />

Days of exposure<br />

Figure 3.--Change of palatability of three clones after extraction. Vertical axis represents the weight of uneaten leaves<br />

<strong>or</strong> extracted residue leaves. Upper and lower graphs each represent the results f<strong>or</strong> one hare.<br />

Effects of Methanol Extracts on Fresh and Residual Leaves<br />

Application of methanol extracts from dried A4 leaves did not affect the palatability of fresh leaves of A4 and N4 but<br />

made the residues after extraction palatable in both clones. In contrast, application of N4 extract made the fresh leaves of A4<br />

unpalatable, but showed no effect on the fresh leaves of N4 n<strong>or</strong> the residual leaves after extraction in both clones.<br />

Palatability Test of Fractions<br />

The filter paper test was not valid f<strong>or</strong> some hares, because they showed little interest in filter paper and ate it only<br />

occasionally. One hare showed constant interest and gave suggestive results. Most of the fractions of A4 extracts showed<br />

high palatability, whereas some of the fractions of N4 extracts showed only weak palatability (Fig. 4). Little of the untreated<br />

filter paper was consumed.<br />

Of A4 extracts, both water-soluble (W) and water-insoluble (M) fractions were highly fav<strong>or</strong>ed. The Wh fraction was<br />

less palatable than tile others. The Mh fraction also showed less palatability than did the Mm. F<strong>or</strong> N4, although the MW<br />

extract was moderately consumed, the separate water-soluble and water-insoluble fractions showed the least consumption.<br />

The water-soluble fractions showed light (Wh) to fairly high palatability (Wb), whereas tile Mh was taken only in small<br />

amounts.<br />

162<br />

i


-


al. 1984, Tahvanainen et al. 1985, Clausen et al. 1986, Hansson et al. 1986, Snyder 1992). Although no specific stimulants<br />

have been rep<strong>or</strong>ted to be involved in differences in palatability between genotypes, the possibility that A4 has some specific<br />

substance(s) that enhance browsing by hares has not been disproved. Our present results only suggest that volatile terpenes<br />

and water-soluble sugars do not have a decisive effect on the differences in the palatability between A4 and other clones.<br />

The A4 clone :isa genotype showing extremely high palatability compared to other plants of the same species. Other<br />

elite clones <strong>or</strong>iginating in Atetsu (Atetsu 1, 3, 5, 6) do not show higher palatability than clones from other areas, hence the<br />

palatability of A4 is difficult to attribute to its geographic <strong>or</strong>igin (Kuwahata and Tanbara 1983). Although many studies<br />

describe large palatability variation within a species, A4 still seems exceptional.<br />

SUMMARY<br />

We examined the chemical fact<strong>or</strong>s determining the extremely high palatability of a sugi clone, A4, to hares. The<br />

palatability of A4 was fairly stable throughout the year. Although A4 had much less essential oil and dichl<strong>or</strong>omethane<br />

extracts than the other clones, steam distillation (<strong>or</strong> boiling) and extraction by n-hexane did not reduce the palatability.<br />

Extraction by methanol made A4 unpalatable, whereas it seemed to make other unpalatable clones m<strong>or</strong>e palatable. Methanol<br />

extracts of A4 had a stimulating effect on feeding whereas that of N4 had a deterrent effect. A4 might contain specific<br />

feeding stimulants that are insoluble in n-hexane, soluble in methanol, and difficult to dissolve in water.<br />

LITERATURE CITED<br />

BRYANT, J.P. 1981. Phytochemical deterrence of snowshoe hare browsing by adventitious shoots of four Alaskan trees.<br />

Science 213: 889-890.<br />

BRYANT, J.P., WIELAND, G.D., REICHARDT, P.B., LEWIS, V.E., and MCCARTHY, M.C. 1983. Pinosylvin methyl<br />

ether deters snowshoe hare feeding on green alder. Science 222: 1023-1025.<br />

CHtBA, S., OGAWA, A., NAGATA, Y., and TOMAKI, K. 1991. Effects of solvent extracts on resistance of Japanese larch<br />

and Kurile larch to voles. Proc. Hokkaido Branch, Japan Wood Res. Soc. 23:70-72 (in Japanese).<br />

CLAUSEN, T.P., REICHARDT, P.B., and BRYANT, P.B. 1986. Pinosylvin and pinosylvin methyl ether as feeding deterrents<br />

in green alder. J. Chem. Ecol. 12:2117-2131.<br />

DIMOCK, E.J., SILEN, R.E., and ALLEN, V.E. 1976. Genetic resistance in Douglas-fir to damage by snowshoe hare and<br />

black-tailed deer. F<strong>or</strong>. Sci. 22: 106-121.<br />

FARENTINOS, R.C., CAPRETTA, P.J., KEPNER, R.E., and LITTLEFIELD, V.M. 1981. Selective herbiv<strong>or</strong>y in tasseleared<br />

squirrels: role of monoterpenes in ponderosa pines chosen as feeding trees. Science 213:1273-1275.<br />

HANSSON, L., GREE R., LUNDREN, L., and THEANDER, O. 1986. Susceptibility to vole attacks due to bark phenols<br />

and terpenes in Pinus cont<strong>or</strong>ta provenances introduced into Sweden. J. Chem. Ecol. 12: 1569-1578.<br />

HARA, M. 1973. The variation of hare damage among sugi clones. Proc. Tree Breed. Soc. 63-66 (in Japanese).<br />

HIRAKAWA, H. 1989. Screening of the preferential feeding substance(s) of the Japanese hare (Lepus brachyurus) included<br />

in the Japanese cedar (Sugi, Cryptomeriajaponica). Proc. Kanto Branch, Jap. Soc. F<strong>or</strong>. 41:145-147 (in Japanese).<br />

KUWAHATA, T. and TANBARA, T. 1983. The variation of susceptibility to hare damage among sugi elite clones. Proc.<br />

Kansai Branch, Jap. F<strong>or</strong>. Soc. 34:213-216 (in Japanese)<br />

MATSUURA, T: and MAEDA, M. 1981. Damage from red-backed vole (Clethrionomys rufocanus bedf<strong>or</strong>diae) to Saghalien<br />

fir (Abies sachalinensis) varies with each strain. F<strong>or</strong>. Tree Breed. Hokkaido 24:15-20 (in Japanese).<br />

164<br />

il


OGAWA, A., NAGATA, Y., and SUKENO, S. 1992. Effects of solvent extraction on resistance of the hybrid between<br />

Japanese larch and Kurile larch to voles. Proc. Hokkaido Branch, Japan Wood Res. Soc. 24:61-64 (in Japanese).<br />

RADWAN, M.A. 1972. Differences between Douglas-fir genotypes in relation to browsing preference by black-tailed deer.<br />

Can. J. F<strong>or</strong>. Res. 2: 250-255.<br />

RADWAN, M.A., CROUCH, G.L., HARRINGTON, C.A., and ELLIS, W.D. 1982. Terpenes of ponderosa pine and feeding<br />

preferences by pocket gophers. J. Chem. Ecol. 8:241-253.<br />

REICHARDT, RB., BRYANT, J.P., CLAUSEN, T.R, and WlELAND, G.D. 1984. Defence of winter-d<strong>or</strong>mant Alaska paper<br />

birch against snowshoe hares. Oecologia 65: 58-69.<br />

ROUSI, M., TAHVANAINEN, J., and UOTILA, I. 1989. Inter- and intraspecific variation in the resistance of winterd<strong>or</strong>mant<br />

birch (Betula spp.) against browsing by the mountain hare. ttolar. Ecol. 12: 187-1192.<br />

ROUSI, M., TAHVANAINEN, J., and UOTILA, I. 1991. A mechanism of resistance to hare browsing in winter-d<strong>or</strong>mant<br />

European white birch (Betula pendula). Amer. Natur. 137: 64-82.<br />

ROUSI, M., TAHVANAINEN, J., and UOTILA, I. 1993. Effects of shading and fertilization on resistance of winter-<br />

d<strong>or</strong>mant birch (Betula pendula)to voles and hares. Ecology 74: 30-38.<br />

SNYDER, M.A. 1992. Selective herbiv<strong>or</strong>y by Abert's squirrel mediated by chemical variability in Ponderosa pine. Ecology<br />

73: 1730-1741.<br />

TAHVANAINEN, J., HELLE, E., JULKUNEN-TIITTO, R., and LAVOLA, A. 1985. Phenolic compounds of willow bark as<br />

deterrents against feeding by mountain hare. Oecologia 65:319-323.<br />

165


GENETIC AND PHENOTYPIC VARIATION IN THE INDUCED REACTION OF<br />

SCOTS PINE TO LEPTOGRAPHIUM WINGFIELDII:<br />

REACTION ZONE LENGTH AND FUNGAL GROWTH<br />

F. LIEUTIER 1,B. LANGSTROM z, H. SOLHEIM 3<br />

C. HELLQVIST 2, and A. YART l<br />

_<strong>Station</strong> de Zoologie F<strong>or</strong>esti_re, Institut National de la Recherche Agronomique, Ardon, 45 160 Olivet, France<br />

2Division of F<strong>or</strong>est Entomology, Swedish University of Agricultural Sciences, S-77073 Garpenberg, Sweden<br />

3Department of F<strong>or</strong>est Ecology, N<strong>or</strong>wegian F<strong>or</strong>est <strong>Research</strong> Institute, As-NLH, N 1432, N<strong>or</strong>way.<br />

INTRODUCTION<br />

The induced defensive reaction of phloem generally plays a decisive role in the resistance of conifers to attack by<br />

bark beetles and their associated fungi. It is thought to be induced by fungi (Reid et al. 1967, Berryman 1972, Raffa and<br />

Berryman 1983, Christiansen and H<strong>or</strong>ntvedt 1983, Matson and Hain 1985, Christiansen et al. 1987, Lfingstr6m et al. 1992),<br />

<strong>or</strong> by the b<strong>or</strong>ing activity of the beetle (Lieutier et al. 1988, Lieutier 1993). In most cases, by stimulating that energy demanding<br />

response, the fungi can lower the threshold of attack density above which the tree is overwhelmed and beetle attacks are<br />

successful. In all situations, the individual reaction is the basic mechanism of tree resistance.<br />

The between-tree variability in the threshold of attack density (Waring and Pitman 1980, Mulock and Christiansen<br />

1986), as well as the between-tree variability in the induced reaction itself (Shrimpton and Reid 1973, Peterman 1977,<br />

Lieutier and Ferrell 1989, Lieutier et al. 1993), has been linked to tree vig<strong>or</strong>. However, the genetic contribution to <strong>this</strong><br />

variability has never been studied.<br />

In Scots pine, Pinus sylvestris L., an induced reaction is essential to resist attacks by lps sexdentatus Boern. and<br />

Tomicus piniperda L., and their associated fungi (Lieutier et al. 1988, 1989a; Solheim and L_ngstr6m, 1991; L_ngstr_m et al.<br />

1992). The blue-staln fungus, Leptographium wing_i'eldii M<strong>or</strong>elet, associated with T. piniperda, does not seem to play a role<br />

in the development of the induced reaction and in the success of attacks (Lieutier et al. 1988, 1989b). However, artificial<br />

inoculation of its sp<strong>or</strong>ulated cultures into Scots pine always incites strong defensive reactions (Lieutier et al. 1989a, Solheim<br />

and L_ngstr6m 1991) and the tree's response depends on the number of sp<strong>or</strong>es present in the inoculum (Lieutier et al. 1989b).<br />

Additionally, it may play a role in tree m<strong>or</strong>tality, since weakened Scots pine can be killed by a high density of artificial<br />

inoculations (Solheim et al. 1993).<br />

The present study is part of a cooperative project with the following objectives: (1) to study the variation in the<br />

induced reactions (reaction zone f<strong>or</strong>mation, sapwood occlusion) between Scots pine provenances in two different climatic<br />

regions, (2) to study the perf<strong>or</strong>mance of a French and a Swedish strain of L. wingfieldii in these provenances under different<br />

climatic conditions, and (3) to study the defensive chemistry involved in the induced reactions, particularly the roles of resin<br />

acids and phenolics, and to search f<strong>or</strong> chemical markers f<strong>or</strong> resistance among these groups of chemicals. This study addresses<br />

the first two issues.<br />

Mattson, W.J., Niemelfi, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle.<br />

<strong>USDA</strong> F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-I83, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

166


MATERIALS AND METHODS<br />

Field W<strong>or</strong>k<br />

In early spring, 1991, provenances of Scots pine were selected in central France (Arb<strong>or</strong>etum des Barres, Nogent-sur-<br />

Vernisson) and in southern Sweden (Remningst<strong>or</strong>p). Eleven (55 trees) and l0 (60 trees) provenances were chosen in France<br />

and Sweden respectively, so as to cover as wide a geographic distribution as possible, while taking into account the necessity<br />

of considering comparable provenances between the 2 countries (Table 1). In addition, the best lodgepole pine, Pinus<br />

cont<strong>or</strong>fa, provenance was selected at Remningst<strong>or</strong>p (6 trees) f<strong>or</strong> a separate comparison with the Scots pines. French trees had<br />

been planted between 1951 and 1957, except f<strong>or</strong> 2 provenances planted in 1941 (Table 1), using 3-year-old seedlings. Their<br />

DBH ranged fiom 14-22.5 cm. Swedish trees had been planted in 1962 using 4-year-old seedlings and their DBH ranged<br />

from 10.5-18 cm.<br />

All trees were inoculated with a French and a Swedish strain of L. winglqeldii, <strong>or</strong>iginally isolated from T. piniperda in<br />

France (Lieutier et al. 1989a) and in Sweden (Solheim and LSngstr6m 1991) respectively. The inoculation materials were 2week-old<br />

cultures on malt agar. Holes were made to cambium depth with a 5-ram c<strong>or</strong>k b<strong>or</strong>er, 5-ram pieces of agar culture<br />

were placed in the holes, and the bark plugs were put back into the tree. In <strong>this</strong> study, we inoculated five rings spaced (ca. 30<br />

cm apart) on the lower stem (:from ca. 60 to 180 cm stern height) in six points spaced evenly around the circumference. Each<br />

of two spots at opposite sides of the stem received sterile agar (control), with French and Swedish strains of L. wing/i'e/dii. In<br />

<strong>or</strong>der to avoid lesions growing into each other, inoculation points in subsequent rings were displaced by 30 degrees relative to<br />

the previous ring. The inoculations took place from 8 to 12April in France and on 26 April in Sweden.<br />

Three weeks after the inoculations, the reaction zones, i.e., the resin-soaked lenticular zones (often referred to as<br />

lesions) surrounding the points of inoculation, were sampled as follows: All 6 inoculation points of ring 2, and one of the two<br />

inoculation points of' each treatment in ring 4 (i.e., 3 inoculation points per treatment in total), were carefully exposed by<br />

removing the outer bark. The reaction zones were measured on the external face of the phloem upwards and downwards<br />

from each inoculation point. F<strong>or</strong> re-isolation of the inoculated fungi, the right half of two reaction zones of each treatment<br />

was cut out, put individually in plastic bags, and cold-st<strong>or</strong>ed immediately at 2°C. The remaining reaction zones were put in<br />

paper bags <strong>or</strong> in vials, immediately frozen on dry ice, and st<strong>or</strong>ed f<strong>or</strong> later chemical analyses. Two unwounded phloem<br />

samples were also collected in each tree and preserved the same way as the remaining reaction zones f<strong>or</strong> chemical analyses.<br />

Six weeks after the inoculations, 3 other inoculation points per treatment, located in rings 3 and 4, were investigated<br />

in each tree and sampled as described above. Two new unwounded phloem samples were also taken. After sampling the<br />

reaction zones, 2 trees in France and 3 trees in Sweden were felled in each provenance, and discs were cut at rings 1,3 and 5.<br />

These discs were st<strong>or</strong>ed in a cold room (2°C) f<strong>or</strong> later examination of sapwood reaction zones and re-isolation of the fungi.<br />

The total height of the felled trees and the lengths of their current leader shoots were measured.<br />

Twenty-one weeks (24 in Sweden) after the inoculations, the reaction zones were sampled (rings I and 5) from the<br />

remaining trees as described above. Trees were felled and discs were cut as described above also.<br />

Diameter and height of the trees differed significantly between provenances in both France and Sweden. In France,<br />

the provenances "Rud" and "Die" had the highest dimensions, while "Spe", "Seg" and "Fin" had the lowest. In Sweden, the<br />

native provenances "Vit", "Vfis" and "Kil" were the tallest and had the biggest diameter, while the provenances "Mal" and<br />

"Enz" were very small and thin. The provenance "Tab" was peculiar, as it had the lowest height although its diameter was<br />

comparable to other provenances. These trees were the last ones left at the plot after snow-break damage. On the contrary,<br />

lodgepole pine was taller, although its diameter did not differ from Scots pine provenances.<br />

In France, no between-provenance differences could be noticed in the length of the current leader shoots, except in<br />

September 1991, but damage by T. pit@erda did not permit any conclusions. In Sweden, significant provenance differences<br />

existed. Shoots of the f<strong>or</strong>eign provenances were generally sh<strong>or</strong>ter than that of the local ones. Lodgepole pine also had long<br />

shoots. Six weeks after inoculations, the trees in France had completed ca. 2/3 of their leader growth, whereas trees in<br />

Sweden had only completed ca. 1/2 of theirs.<br />

167


Table 1.--Characteristics of the Scots pine provenances and of P. cont<strong>or</strong>ta studied in Sweden and in France.<br />

Designation <strong>or</strong>igin latitude longitude altitude age<br />

(m) (years)<br />

Proven ances studied in France (Nogent-sur-Vernisson):<br />

Fin (Finland) 40<br />

Kal Kalmar (Sweden) 56° 30' 16° 20' 40 37<br />

Spe Spey valley (U.K., 57° 20' 03° 30'W 41<br />

Scotland)<br />

Rud Rudczany (Poland) 53° 40' 21° 29' 130 53<br />

Klo Klosterreichenbach 48° 30' 08° 25' 43<br />

(Germany, Schwarzwald)<br />

Niz Saint Nizier de F<strong>or</strong>nas 45° 25' 04° 05' 900 42<br />

(France, Loire)<br />

Die Saint Di6 (France, 48° 15' 07 ° 400 43<br />

Vosges)<br />

Mat La Matte des Angles 42" 36' 02° 07' 1520 53<br />

(France, Pyr_ndes)<br />

Ser (Serbie) 43 ° 21° 37<br />

Seg Segovie (Spain) 40° 52' 04°W 1400 41<br />

Can Canakkale (Turkey) 40° 10' 26° 25' 37<br />

Provenances studied in Sweden (Remningst<strong>or</strong>p):<br />

168<br />

Mal Malfi (Sweden) 65° 38' 21° 07' 75 33<br />

Sve Sveg (Sweden) 62" 03' 14° 19' 385 33<br />

Kil Kilaf<strong>or</strong>s (Sweden) 61° 08' 16° 31' 150 33<br />

Vat Varnhem (Sweden) 58° 22' [3 ° l l' 270 33<br />

V_is V_istervik (Sweden) 57° 46' 16° 39' 20 33<br />

Vit Vittsk6vle (Sweden) 55° 51' 14° 01' 30 33<br />

O1o Olonets (Russia, 61° 33° 1O0 33<br />

Carelia)<br />

]nv Inverness (U.K., 57° 10' 05°W 300 33<br />

Scotland)<br />

Tab Tab<strong>or</strong>z (Poland) 53° 45' 20° 06' 130 33<br />

Enz Enzkl6sterle 48* 45' 08° 30' 700 33<br />

(German_, Schwarzwald)<br />

Con Stuart lake (Canada, 54* 30' 124° 15'W 600 33<br />

B.C.) = P. cont<strong>or</strong>ta<br />

i


Lab<strong>or</strong>at<strong>or</strong>y Procedures<br />

The week after collection, the samples were analyzed f<strong>or</strong> fungus extension. From the reaction zones, phloem pieces<br />

were cut with a sterile raz<strong>or</strong> blade at 5 mm intervals upwards and downwards from the inoculation points beyond the visible<br />

reaction zone. The sample pieces were placed on petri dishes with malt agar medium and incubated at 25°C f<strong>or</strong> 1 week.<br />

Positive and negative rec<strong>or</strong>ds of L. wingfieldii were noted. Sapwood samples were handled the same way, except that<br />

sampling took place in the radial direction. The maximum extension of the sapwood reaction zone was also measured<br />

radially.<br />

Data Handling and Statistics<br />

All caiculations were done with the SAS statistical package. Preliminary analyses of vertical reaction zone lengths<br />

and fungal growth revealed no differences between the vertical extension of reaction <strong>or</strong> fungal zones upwards and downwards<br />

from the inoculation points regardless of the treatment (Student's t-test). Neither were there any differences between<br />

reaction <strong>or</strong> fungal zones from different inoculation rings of the same treatment. Hence, we pooled data within the same tree,<br />

and used tree-wise treatment means (based on 3 lesions <strong>or</strong> less in some cases with missing data) as calculation units.<br />

Treatment means were tested using one-way analysis of variance followed by Tukey's test f<strong>or</strong> multiple comparisons.<br />

Two-way analysis of variance was also employed, as were pair-wise and standard t-tests, depending on the situation. Interdependency<br />

between different variables was expl<strong>or</strong>ed by regression analyses.<br />

Reaction Zone Length<br />

Phloem<br />

Total length of the reaction zone varied significantly between provenances and between treatments (Fig. 1). In<br />

Sweden, local provenances generally exhibited longer reaction zones than f<strong>or</strong>eign provenances at 6 and 24 weeks. The<br />

opposite tendency however, was visible in France at 6 weeks, where the longest reaction zones were observed in the southern<br />

f<strong>or</strong>eign provenances, particularly the Spanish and Serbian provenances, while the sh<strong>or</strong>test reaction zones belonged to the<br />

local provenances. At Remningst<strong>or</strong>p, lodgepole pine exhibited sh<strong>or</strong>ter reaction zones than all provenances of Scots pine after<br />

3 and6 weeks. Despite existing differences between provenances, no systematic trends (e.g., latitudinal) in reaction zone<br />

lengths could be detected.<br />

At Remningst<strong>or</strong>p, reactions to the two strains of L. wingfieldii were similar but, at Nogent-sur-Vernisson, the French<br />

strain caused a m<strong>or</strong>e extensive reaction zone than did the Swedish strain (Fig. 1 and 2). In both countries, reaction zones<br />

resulting from fungal inoculations were longer than reaction zones to the controls, but the latter were clearly longer in France<br />

than in Sweden. The c<strong>or</strong>relation of the reaction zone length to the fungus increased m<strong>or</strong>e rapidly and to a higher level in<br />

Sweden than in France (Fig. 2). A plateau was generally reached after 6 weeks, except f<strong>or</strong> the French fungus strain in France<br />

where the reaction zones still expanded after 6 weeks. A few provenances, however, did not fit with that description (Fig. 1).<br />

Few of the provenances studied occurred at both study sites. The pair "Kal" (France) and "Vas" (Sweden) <strong>or</strong>iginates<br />

from the same region and altitude in southeastern Sweden, and should hence be comparable. We can also assume that the<br />

Finnish "Fin" (France) and Carelian "O1o" (Sweden) as well as the pairs from Scotland, Poland, and Germany are fairly<br />

similar. However, a closer examination of these pairs of provenances does not disclose any deviations from the general<br />

results given above.<br />

Fungal Growth<br />

Total vertical fungal expansion varied significantly between provenances only f<strong>or</strong> the 6 week sampling date (Fig. 3).<br />

At Nogent-sur-Vernisson, fungal growth was the most expanded in the southern, f<strong>or</strong>eign provenance "Seg", and the least<br />

expanded in the French provenances, particularly the Pyrenean one. At Remningst<strong>or</strong>p, fungal growth was the most expanded<br />

169


Figure l.--Vertical extension of induced reaction zones in Scots pine phloem in response to artificial inoculations with a<br />

French (black) and Swedish (dotted) strain of Leptographium wingfieldii as well as a sterile control (white), in an<br />

experhnent in France (left) and Sweden (right) using different provenances of Scots pine and one of lodgepole pine<br />

(Sweden only). Samples were taken after 3 (A), 6 (B) and 21 (C) (France) <strong>or</strong> 24 (C) (Sweden) weeks; provenance<br />

abbreviations are explained in Table 1. Numbers of observations (trees) were 5 and 6 (France and Sweden, respectively)<br />

f<strong>or</strong> the first two sampling dates, and 3 f<strong>or</strong> last sampling date at both sites. Vertical bars indicate standard err<strong>or</strong>.<br />

P-values refer to two-way-anovas f<strong>or</strong> differences between treatments and provenances.<br />

17(i)


E<br />

120<br />

Nogent-sur-Vernisson Remningst<strong>or</strong>p<br />

c<br />

occ_ E. 80 11"_ i<br />

40<br />

E 20<br />

ot,,-<br />

E_<br />

0 ......... _...........<br />

0 6 12 18 24 0 6 12 18 24<br />

30 ..... • .....<br />

C_10 - _<br />

C<br />

I1.<br />

0 0 6 12 18 24 0 6 12 18 24<br />

Weeks after inoculation Weeks after inoculation<br />

Figure 2.--Vertical progress in reaction zone size (upper row) and fungal growth (lower row) in Scots pine phloem in<br />

sterile control (triangle), in an experiment in France (left) and Sweden (right). Tree provenances were pooled.<br />

response to artificial inoculation with a French (circle) and Swedish (square) strain ofL. wingfieldii as well as a<br />

Samples were taken after 3, 6 and 21 (France) <strong>or</strong> 24 (Sweden) weeks. Numbers of observations (trees) were 5 and 6<br />

(France and Sweden, respectively) f<strong>or</strong> the first two sampling dates, and 3 f<strong>or</strong> last sampling date at both sites. Vertical<br />

bars indicate standard err<strong>or</strong>.<br />

in the native provenance "Ki!". No significant differences were observed between treatments, except f<strong>or</strong> the 21 week<br />

sampling date at Nogent-sur-Vernisson, where the French strain of L. wingfieldii grew further than the Swedish strain.<br />

However, in the Spanish provenance, that was also true f<strong>or</strong> all sampling dates. In both locations, fungi expanded rapidly<br />

during the first 3 weeks, and thereafter stabilized (Fig. 2). At Nogent-sur-Vernisson, fungus expansion was higher f<strong>or</strong> the<br />

French than f<strong>or</strong> the Swedish strain after 21 weeks. The expansion of the Swedish strain even decreased at 21 weeks, except<br />

in the Pyrenean provenance (Fig. 3).<br />

17l


E<br />

Nogent-sur-Vernisson Remningst<strong>or</strong>p<br />

75 .....................................................................................................................................<br />

E_ 50<br />

E<br />

i A Ptreatm= 0.1948 A Ptreatm = 0.6333<br />

r t t<br />

Pprov = 0.1211: Pprov = 0.0817<br />

0 ....................................................<br />

75 .............................................................................................................................................<br />

------1 ..............................<br />

B Ptreatm= 0.1030 1 B Ptreatm= 0.9165<br />

i.+:<br />

o I t::i::i<br />

I<br />

E 50, Pprov = i0-0399_1 Pprov = 0.0099<br />

_'_2si , i<br />

It.<br />

]5 r.........................................................................................................................................................................................<br />

C Ptreatm= 0.0027 1 mal sve kil var vs vit olo inv tab enz con<br />

0.3669<br />

1<br />

E , I"Pr°V<br />

E , --I<br />

0 ...........L ...........................................<br />

] ]-<br />

fin kal spe rud klo niz die mat ser seg can<br />

Figure 3.---Vertical expansion of fungal growth after inoculation of a French and Swedish strain of L. wingfieldii in the<br />

172<br />

phloem of different pine provenances in an experiment in France (left) and Sweden (right). Samples were taken after<br />

3 (A), 6 (B) and 21 (C) (France only) weeks. Provenance abbreviations are explained in Table 1. Numbers of<br />

observations (trees) were 5 and 6 (France and Sweden, respectively) f<strong>or</strong> the first two sampling dates, and 3 fk)r last<br />

sampling date (France only). Vertical bars indicate standard err<strong>or</strong>. P-values refer to two-way-anovas f<strong>or</strong> differences<br />

between treatments and provenances.


Interdependence Between Reaction Zone Length and Fungal Growth<br />

In all cases, fungat growth was much less expansive than reaction zone length (cf. Fig. l - 3). These two parameters<br />

were positively c<strong>or</strong>related with each other after 6 weeks (r=0.73 and 0.51 in France and Sweden, respectively, and after 21<br />

weeks (r=0.71, France only), but not after 3 weeks (r=0.38 and -0.27 in France and Sweden, respectively).<br />

Radial Occlusion and Fungal Growth Into Sapwood<br />

As sapwood occlusion was minimal (Fig. 4), data f<strong>or</strong> the different provenances are not given in detail. After 6 weeks,<br />

the range in sapwood occlusion means was 0.4-2.1 mm and 0.2-3.1 mm in France and Sweden, respectively. After 21/24<br />

weeks, the c<strong>or</strong>responding ranges were 1.1-5.8 mm and 0.7-4.7 ram. Sapwood occlusion was always m<strong>or</strong>e developed after<br />

fungus inoculation than after control inoculation, and continued to expand between 6 and 21/24 weeks (Fig. 4). This was true<br />

f<strong>or</strong> both localities. At Nogent-sur-Vernisson, the French strain of L. wingfieldii caused a larger sapwood occlusion than did<br />

the Swedish strain, 21 weeks after inoculation.<br />

,5<br />

4<br />

3<br />

2<br />

1<br />

OE<br />

4<br />

3<br />

Nogent-su r-Vernisson Remningst<strong>or</strong>p<br />

.._..t--& A<br />

o 0 6 12 18 24 0 6 12 18 24<br />

Weeksafter inoculation Weeksafter inoculation<br />

Figure 4._Radial occlusion of sapwood (upper row) and radial expansion of fungi (lower row) following inoculation with a<br />

French (circle) and Swedish (square) strain of L. wingfieldii as well as a sterile control (triangle), in different Scots<br />

pine provenances, in an experiment in France (left) and Sweden (right). Samples were taken after 6 and 21 weeks<br />

(France) <strong>or</strong> 6 and 24 weeks (Sweden). Provenance abbreviations are explained in Table 1. Numbers of observations<br />

(trees) were 2 and 3 (France and Sweden, respectively) f<strong>or</strong> the 6 week sample, and 3 f<strong>or</strong> the last sampling date (both<br />

sites). Vertical bars indicate standard err<strong>or</strong>. 173<br />

T


As f<strong>or</strong> radial occlusion, fungal growth in the sapwood was minimal (Fig. 4), and did not differ between provenances.<br />

Hence, data are not given in detail f<strong>or</strong> the different provenances. After 6 weeks, the average fungal penetration into the<br />

sapwood ranged from 0.2-2.0 mm and from 0.2-6.0 mm in France and Sweden, respectively. After 21/24 weeks, the c<strong>or</strong>responding<br />

ranges were 0.0-1.6 mm and 0.0-9.0 ram. In both countries, fungi generally regressed <strong>or</strong> remained stable from 6 to<br />

21/24 weeks (Fig. 4). There were no differences between the 2 strains of fungi, but both fungi expanded further into the<br />

sapwood in Sweden than in France.<br />

DISCUSSION<br />

Comparison Between Provenances<br />

The observed differences between provenances in all variables studied (reaction zone length, sapwood occlusion, and<br />

fungal growth - vertically as well as radially) do not demonstrate a systematic pattern justifying general conclusions about<br />

provenance differences. In Sweden however, f<strong>or</strong>eign provenances responded differently than native provenances to fungus<br />

infection. As <strong>this</strong> was not the case in France, no latitudinal <strong>or</strong> altitudinal trends can be detected. On the other hand, systematic<br />

differences in defense reactions may have been masked by the considerable inter-tree variation, possibly due to variations<br />

in micro-ecological conditions from one provenance plot to another, <strong>or</strong> even to genetic differences between trees of the same<br />

provenance. A larger number of replications would have been required. Even better would have been to w<strong>or</strong>k with clones<br />

instead of provenances. Thus in <strong>this</strong> experiment, reaction zone size alone did not provide enough intbrmation to evaluate<br />

possible genetic differences in host resistance to blue-stain fungi.<br />

In Sweden, lodgepole pine also responded in the same manner as f<strong>or</strong>eign Scots pine provenances, which suggests it<br />

may be similarly susceptible to European bark beetles.<br />

Comparison Between Treatments<br />

As observed earlier, fungal inoculations produced larger reaction zones than the sterile zones (Wong and Berryman<br />

1977, Solheim 1988, Ross et al. 1992). F<strong>or</strong> the variables measured, the two fungal strains caused larger differences in France<br />

than in Sweden, and <strong>this</strong> locality effect will be discussed below. The French strain of L. wingfieldii invariably caused larger<br />

reaction zones than the Swedish strain in France, and similar reaction zones in Sweden. As the pattern was similar f<strong>or</strong> fungat<br />

growth, the French strain could be considered m<strong>or</strong>e aggressive than the Swedish one. The observed difference cannot be<br />

attributed to some environmental <strong>or</strong> host fact<strong>or</strong>, as inoculations of the two strains occurred parallel in the same trees. Thus,<br />

some inherent difference between the fungal strains should cause <strong>this</strong> difference in aggressiveness. F<strong>or</strong> sapwood, the French<br />

strain occluded m<strong>or</strong>e tissue than the Swedish strain after 21 weeks in France, hence its pathogenicity also could be higher. In<br />

all treatments however, sapwood occlusion was shallow, despite extensive reaction zone f<strong>or</strong>mation in the phloem (cf.<br />

Parmeter et al. 1992). This can be explained by the low density of inoculation, allowing the tree to contain the infection.<br />

Comparison Between Localities<br />

All variables measured indicated clear differences between localities. Reactions zones following fungal inoculations<br />

were larger in Sweden than in France, as was sapwood occlusion. That was also the case f<strong>or</strong> fungal growth, although not so<br />

clearly tbr the phloem. In contrast, sterile inoculations caused larger reaction zones in France than in Sweden f<strong>or</strong> the phloem,<br />

but not f<strong>or</strong> the sapwood.<br />

As these differences were also true inside comparable pairs of provenances (e.g., "Kal" vs "Vas", "Fin" vs "Olo",<br />

Spe" vs "Inv", "Rud" vs "Tab" and "Klo" vs "Enz"), a provenance effect can be discarded. However, the finding that sterile<br />

inoculations resulted in larger reaction zones in France than in Sweden whereas fungal inoculations yielded the opposite<br />

pattern, is difficult to explain. This contradiction implies a higher sensitivity to mechanical wounding in France than in<br />

Sweden, but also a m<strong>or</strong>e efficient containment of the fungi in the f<strong>or</strong>mer than in the latter case. Possibly, differences in<br />

prevailing temperature conditions may have fav<strong>or</strong>ed fungal growth in Sweden, as L. wingfieldii is known to have a low<br />

temperature preference (Lieutier and Yart 1989).<br />

Although the experimental conditions were planned to be as similar as possible, trees may have differed in phenology<br />

between the localities. Judging from the relative shoot lengths 6 weeks after inoculation, L<strong>or</strong>io's growth-differentiation<br />

174<br />

:i


alance hypothesis (L<strong>or</strong>io 1986) cannot explain the locality effect on trees. Indeed, trees at Remningst<strong>or</strong>p were in a m<strong>or</strong>e<br />

intensive phase of growth than the trees at Nogent-sur-Vernisson, and should have had larger reaction zones in response to<br />

sterile inoculation. Age of the trees may be a m<strong>or</strong>e useful explanation, as trees were clearly older in France than in Sweden,<br />

and it has already been rep<strong>or</strong>ted that older Scots pines have larger reaction zones than younger ones (Lieutier and Ferrell<br />

1989, Lieutier et hi. 1993).<br />

Although not demonstrated in our experiment,<br />

General<br />

genetic<br />

Conclusions<br />

variability in the tree's response to invading fungi (<strong>or</strong> beetles)<br />

probably exists in Scots pine. In addition, we have demonstrated that there is an inherent variability in strains of L.<br />

win_ieldii. However, phenotypic (geographic) variations concerning both tree and fungus, are superimposed on <strong>this</strong> pattern.<br />

This can be seen in the situation with larger reaction zones to fungus in Sweden than in France, whereas the opposite pattern<br />

occurredtions, f<strong>or</strong> the control. One must keep in mind all these potential variations while studying tree/bark-beetle/fungus interac-<br />

SUMMARY<br />

The aggressiveness and pathogenicity of one French and one Swedish isolate of Leptographium wingfieldii were<br />

studied in different Scots pine provenances, both in France and Sweden. After inoculations, reaction zone length and fungal<br />

growth varied significantly between provenances, at least in the phloem, but no consistent trends could be detected. In<br />

France, fungal growth was m<strong>or</strong>e extended and the reaction zones were m<strong>or</strong>e developed f<strong>or</strong> the French than f<strong>or</strong> the Swedish<br />

strain, both in phloem and sapwood. In Sweden, fungal growth and reaction zone length were comparable f<strong>or</strong> the two strains.<br />

Fungal growth and reaction zones induced by fungus inoculations were m<strong>or</strong>e extensive in Sweden than in France, but<br />

reaction zones induced by sterile wounds were larger in France. Fungus inoculations ' however, always induced larger<br />

reaction zones than sterile inoculations.<br />

ACKNOWLEDGEMENTS<br />

This study is part of ongoing cooperative research between Institut National de la Recherche Agronomique (INRA),<br />

Swedish University of agricultural Sciences (SLU), and N<strong>or</strong>wegian F<strong>or</strong>est <strong>Research</strong> Institute (NISK), within the field of bark<br />

beetle/blue stain fungi/host tree interactions. A grant from the "Hildur and Sven Wingqvist's foundation f<strong>or</strong> f<strong>or</strong>est research"<br />

facilitated the Swedish part of the study, and is gratefully acknowledged. The CEMAGREF and the Arb<strong>or</strong>etum des Barres at<br />

Nogent-sur-Vernisson are also acknowledged f<strong>or</strong> providing experimental plots f<strong>or</strong> the French part of the study. The auth<strong>or</strong>s<br />

express their gratitude to J. Garcia and P. Romary from INRA and to H. Knutsen, C. Nelleman and O. Olsen from NISK f<strong>or</strong><br />

their technical help.<br />

LITERATURE CITED<br />

BERRYMAN, A.A. 1972. Resistance to bark beetle-fungi associations. BioScience 22: 598-602.<br />

CHRISTIANSEN, E. and HORNTVEDT, R. 1983. Combined Ips-Ceratocystis attack on N<strong>or</strong>way spruce, and defensive<br />

mechanisms of the trees. Z. angew. Entomol. 96:110-118.<br />

CHRISTIANSEN, E., WARING, R.H., and BERRYMAN, A.A. 1987. Resistance of conifers to bark beetle attack: searching<br />

f<strong>or</strong> general relationships. F<strong>or</strong>. Ecol. Manage. 22: 89-106.<br />

L]kNGSTROM, B., HELLQVIST, C., ERICSSON, A., and GREF, R. 1992. Induced defence reaction of Scots pine following<br />

stem attacks by Tomicus piniperda. Ecography 15:318-327.<br />

LIEUTIER, F. 1993. Induced defense reaction of conifers to bark beetles and their associated Ophiostoma species, p. 206-<br />

215. In Wingfield, M.J., Seifert, K.A., and Webber, J.E, eds. Ceratocystis and Ophiostoma: Taxonomy, Biology and<br />

Pathogenicity. Amer. Phytopathol. Soc. Press, St. Paul.<br />

175


LIEUTIER, E and FERRELL, G.T. 1989. Relationships between indexes of tree vigour and the induced defense reaction of<br />

Scots pine to a fungus associated with Ips sexdentatus (Coleoptera: Scolytidae), p. 163-178. In Payne, T.L. and<br />

Saarenmaa, H., eds. Integrated Control of Scolytid Bark Beetles. Virginia Polytech. Inst. St. Univ.<br />

LIEUTIER, E, YART, A., GARCIA, J. POUPINEL, B., and Lt_VIEUX, J. 1988. Do fungi influence the establishment of<br />

bark beetles in Scots pine? p. 317-330. In Mattson, W.J., Ldvieux, J., and Bernard-Dagan, C., eds. Mechanisms of<br />

Woody Plant Defenses Against Insects: Search f<strong>or</strong> Pattern. Springer, New Y<strong>or</strong>k.<br />

LIEUTIER, E, YART, A., GARCIA, J., HAM, M.C., MORELET, M., and Lt_VIEUX, J. 1989a. Champignons<br />

phytopathogbnes associds _ Ips sexdentatus Boern. et Tomicus piniperda L. (Coleoptera: Scolytidae) et dtude<br />

prdliminaire de leur agressivitd pour le pin sylvestre. Ann. Sc. F<strong>or</strong>. 46: 201-216.<br />

LIEUTIER, F., CHENICLET, C., and GARCIA, J. 1989b. Comparison of the defense reaction of Pinus pinaster and Pinus<br />

sylvestris to attacks by two bark beetles (Coleoptera: Scolytidae) and their associated fungi. Environ. Entomol. 18:<br />

228-234.<br />

LIEUTIER, E, GARCIA, J., ROMARY, R, YART, A., JACTEL, H., and SAUVARD D. 1993. Inter-tree variability in the<br />

induced defense reaction of Scots pine to single inoculations by Ophiostoma brunneo-ciliatum, a bark beetle associated<br />

fungus. F<strong>or</strong>. Ecol. Manage. 59: 257-270.<br />

LIEUTIER, F. and YART, A. 1989. Preferenda thermiques des champignons associds _ lps sexdentatus Boern. et Tomicus<br />

piniperda L. (Coleoptera: Scolytidae). Ann. Sc. F<strong>or</strong>. 46:411-415.<br />

LORIO, Jr. RL. 1986. Growth-differentiation balance: a basis f<strong>or</strong> understanding southern pine beetle-tree interactions. F<strong>or</strong>.<br />

Ecol. Manage. 14: 259-273.<br />

MATSON, RA. and HAIN, F.R 1985. Host conifer defense strategies: a hypothesis, p. 33-42. In Safranyik, L. and<br />

Berryman, A.A. eds. The role of the host in the population dynamics of f<strong>or</strong>est insects. Proc. IUFRO Conf. Banff<br />

Alberta Canada.<br />

MULOCK R and CHRISTIANSEN E. 1986. The threshold of successful attack by lps O,pographus on Picea abies: a field<br />

experiment. F<strong>or</strong>. Ecol. Manage. 14: 125-132.<br />

PARMETER, J.R., SLAUGHTER, G.W., CHEN, M., and WOOD, D.L. 1992. Rate and depth of sapwood occlusion<br />

following inoculation of pines with blue stain fungi. F<strong>or</strong>. Sci. 38: 34-44.<br />

PETERMAN, R.M. 1977. An evaluation of the fungal inoculation method of determining the resistance of lodgepole pine to<br />

mountain pine beetle (Coleoptera: Scolytidae) attacks. Can. Entomol. 109: 113-448.<br />

RAFFA, K.E and Berryman, A.A. 1983. The role of host resistance in the colonization behavi<strong>or</strong> and ecology of bark beetles<br />

(Coleoptera: Scolytidae). Ecol. Monogr. 53: 27-49.<br />

REID, R.W., WHITNEY, H.S., and WATSON, J.A. 1967. Reactions of lodgepole pine to attack by Dendroctonus<br />

pseudotsugae Hopkins and blue stain fungi. Can. J. Bot. 45:I115-1126.<br />

ROSS, D.W., FENN, R, and STEPHEN, F.M. 1992. Growth of southern pine beetle associated fungi in relation to the<br />

induced wound response in loblolly pine. Can. J. F<strong>or</strong>. Res. 22: 1851-1859.<br />

SOLHEIM, H. 1988. Pathogenicity of some lps typographus-associated blue-stain fungi to N<strong>or</strong>way Spruce. Medd. N<strong>or</strong>.<br />

Inst. Skogf<strong>or</strong>sk. 40:1 - l 1.<br />

SOLHEIM, H. and L]kNGSTROM, B. 199I. Blue-stain fungi associated with Tomicus piniperda in Sweden and preliminary<br />

observations on their pathogenicity. Ann. Sc. F<strong>or</strong>. 48: 149-156.<br />

176


SOLHEIM, H., LANGSTROM, B., and HELLQVIST, C. 1993. Pathogenicity of the blue-stain fungi Leptographium<br />

wingfieldii and Ophiostoma minus to Scots pine: effect of tree pruning and inoculum density. Can. J. F<strong>or</strong>. Res. 23"<br />

1438-1443.<br />

SHRIMPTON, D.M. and REID, R.W. 1973. Change in resistance of lodgepole pine to mountain pine beetle between 1965<br />

and 1973 Can. J. F<strong>or</strong>. Res. 3:430-432<br />

•<br />

WARING, R.H. and PITMAN, G.B. 1980. A simple model of host resistance to bark beetles. Oreg. State Univ. f<strong>or</strong>. Res.<br />

Lab. Res. Note 65. 2 p.<br />

WONG, B.L. and BERRYMAN, A.A. 1977. Host resistance to the fir engraver beetle. 3. Lesion development and contain<br />

ment of infection by resistant Abies grandis inoculated with Trichosp<strong>or</strong>ium symbioticum. Can. J. Bot. 55: 2358-<br />

2365.<br />

177


CAN PHLOEM PHENOLS BE USED AS MARKERS OF SCOTS PINE<br />

RESISTANCE TO BARK BEETLES?<br />

F. LIEUTIER _,F. BRIGNOLAS l, V. PICRON t, A. YART l, and C. BASTIEN 2<br />

t<strong>Station</strong> de Zoologie F<strong>or</strong>esti_re, Institut National de la Recherche Agronomique<br />

Ardon, 45 160 Olivet, France<br />

2<strong>Station</strong> d'Am61i<strong>or</strong>ation des Arbres F<strong>or</strong>estiers, Institut National de la Recherche Agronomique<br />

Ardon, 45 160 Olivet, France<br />

INTRODUCTION<br />

Bark beetle damage is considerable in all coniferous f<strong>or</strong>ests in the temperate zone. In most cases, it results from<br />

beetle attacks surpassing the critical density (numbers/m 2)below which trees can successfully reject them. Above <strong>this</strong><br />

threshold, natural mechanisms of tree resistance are overcome by the beetles and fungi which become successfully established.<br />

The level of <strong>this</strong> critical threshold of attack density is a fundamental expression of the strength <strong>or</strong> capacity of tree<br />

resistance (Berryman 1978, 1982; Christiansen et al. 1987).<br />

The basic phenomenon involved in the resistance is a phloem reaction induced by the invaders and localized around<br />

them. It consists of a rapid synthesis of chemicals, including secondary metabolites such as terpenes and phenols, which<br />

impregnate tree tissues (Reid et al. 1967, Berryman 1972, Shrimpton 1973, Christiansen and H<strong>or</strong>ntvedt 1983, Christiansen et<br />

al. 1987, Lieutier et al. 1988, Lfmgstr6m et al. 1992, Lieutier 1993). The critical threshold of attack density depends on the<br />

quantity of energy the tree is able to rapidly mobilize f<strong>or</strong> various syntheses in the reaction zones, and <strong>this</strong> quantity is necessarily<br />

limited (Christiansen et al. 1987). Thus, the threshold can be high if the lag time f<strong>or</strong> stopping an aggress<strong>or</strong> at each<br />

point of attack is very sh<strong>or</strong>t and hence the energy mobilized f<strong>or</strong> that purpose in each reaction zone is low. Under these<br />

conditions of arrested attacks, a sh<strong>or</strong>t reaction zone has been assumed to be m<strong>or</strong>e effective than an extended one (Lieutier et<br />

al. 1993).<br />

Markers of conifer resistance to bark beetles could be very useful; e.g., f<strong>or</strong>ecasting susceptibility to beetle attacks and<br />

beetle outbreaks, and aiding in selection of highly resistant trees. Because the concentrations of secondary metabolites vary<br />

considerably during tree responses to aggression, they merit a special interest. Preliminary studies on phloem phenolics have<br />

suggested dramatic changes in Scots pine, Pinus sylvestris L., in response to aggression (Lieutier et al. 1991a). This paper<br />

presents the results from two complementary approaches recently developed in France concerning the possible role of<br />

phenols in arresting bark beetle invasion and their use as biochemical markers of resistance. Detailed methods and other<br />

findings are presented elsewhere (Lieutier et al. 1995).<br />

METHODS<br />

Experimental Devices<br />

In our first experiment, six clones located in the INRA's nursery in Orlfians (France), each represented by one to three<br />

10-year-old trees, were used to test if a relationship existed between reaction zone length and phenolic composition. Trees<br />

were each inoculated with a malt agar sp<strong>or</strong>ulated culture of Ophiostoma brunneo-ciliatum Math.-K., a fungus previously<br />

isolated from lps sexdentatus Boern., and thought to play a role in the establishment of <strong>this</strong> beetle in Scots pine (Lieutier et<br />

al. 1988). Each tree received three inoculations. Three weeks later, the reaction zone around each inoculation point was<br />

measured and sampled. Three samples of unwounded phloem were also taken from each tree.<br />

Mattson, W.J., Niemel_i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle.<br />

<strong>USDA</strong> F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

178


In a second experiment, three 30-year-old Scots pines located in the Orldans f<strong>or</strong>est were inoculated with O. brunneo-<br />

ciliatum to study the kinetics of the trees' phenolic responses. Each tree received 33 inoculations distributed in 5 rings, each<br />

separated by 35 cm. After 3, 7, 14, 30 and 60 days, the reaction zones around respectively 15, 7, 5, 3 and 3 inoculation points<br />

Inoculations were always perf<strong>or</strong>med with 5 mm diameter discs of culture placed in cambium-deep, equally sized<br />

holes<br />

per tree<br />

made<br />

were<br />

with<br />

measured<br />

a c<strong>or</strong>k b<strong>or</strong>er,<br />

and sampled.<br />

and the<br />

At<br />

bark<br />

days<br />

plugs<br />

30<br />

were<br />

and 60,<br />

put<br />

samples<br />

back into<br />

of<br />

the<br />

unwounded<br />

holes, hnmediately<br />

phloem were<br />

after<br />

also<br />

being<br />

collected<br />

collected<br />

from<br />

from<br />

each<br />

the<br />

tree.<br />

trees, the phloem samples were frozen in dry-ice. In the lab<strong>or</strong>at<strong>or</strong>y, all samples were freeze-dried.<br />

second experiment, zone<br />

Phenols Extraction and Analysis<br />

3<br />

which were pooled all together (day 3) <strong>or</strong> by groups of 1 to 3 samples (day 7) to have sufficiently a high<br />

quantity of material f<strong>or</strong> further chemical analyses. Extractions were perf<strong>or</strong>med at 4°C on ground samples. In <strong>or</strong>der to<br />

remove resinous compounds (Alcubilla 1970), phloem powder was first washed with pentane bef<strong>or</strong>e extraction of phenolics<br />

using 80% methanol. Pentane treatment was checked to have no significant effect on the compounds analyzed in our study.<br />

Gallic acid was used as internal standards f<strong>or</strong> HPLC analyses.<br />

Analyses were perf<strong>or</strong>med with reversed HPLC (Waters 600E, Photodiode array detect<strong>or</strong> 991), in a 250 mm column<br />

and 4 mm internal diameter. <strong>Station</strong>ary phase was a C-18 grafted silica (Merck, Lichrospher RRP18) with 0.005 mm p<strong>or</strong>osity.<br />

Mobile phase was a mixture of a 1% acetic acid in ultra pure water (solvent A) with a solution of pure methanol, acetonitrile,<br />

acetic acid (49.5:49.5:1) (solvent B). The gradient is represented in Fig. 1. Readings were made at 310 nm. Results were<br />

expressed in internal standard equivalents per gram of freeze-dried powder.<br />

"" 100<br />

o,_<br />

.¢<br />

= _?IP.,l<br />

0.6 ....................... _ .....................................................<br />

c_<br />

ca ca gll<br />

ca ' 50<br />

0 4 ..... "_ = ........................ ° _I ................. ; ..... "'- s ................................<br />

_ "_<br />

• _I ca . o _<br />

c_ = ...... -._<br />

:,.7,, ; 6 =<br />

t •_ $" .."<br />

_' "<br />

= __ li !_ -r,<br />

.... :_ .......... i................. ! .................. i_ ................ _"' = ........ i........ u<br />

: 2 : "_ : "=, !<br />

02-, . _:.<br />

,v,,,,,. " '" I .... I " " " " I " '-" ' r • _"i'" I " " " " I " " "'"" I ...... ; ' i'''' " " " " " " _<br />

0 20 40 60 80 100 (ran)<br />

time (ram)<br />

Figure 1._Chromatogram of a reaction zone in Scots pine phloem and HPLC gradient. Solvent A = acetic acid, water<br />

(1/99); Solvent B = methyl alcohol, acetonitril, acetic acid (49.5/49.5/1).<br />

179


Characterization of compounds was made by using various methods: two-dimensional thin layer chromatography<br />

(TLC) on cellulose plates to observe flu<strong>or</strong>escence and to test specific chemical reagents, observation of spectral characteristics<br />

in U.V., cochromatography in HPLC with spots obtained from TLC and with various standards. The presence of glycosidic<br />

<strong>or</strong> ester links was tested with acid <strong>or</strong> alkaline hydrolysis, respectively.<br />

RESULTS AND DISCUSSION<br />

Comparison Between Unwounded and Inoculated Phloem<br />

Fourteen peaks were observed, of which eight were characterized at least at the chemical phenol family level (Fig. 1).<br />

Considerable differences existed between unwounded and inoculated phloem (Fig. 2). The most striking concerned the<br />

appearance of stilbenes, pinosylvin (Ps) and its monomethylether (Psme), and of one flavonoid, pinocembrin (Pc), in the<br />

reaction zone of all trees, while these compounds were always totally absent in the unwounded phloem. In all trees, the<br />

concentration of all compounds of the hydroxycinnamic acid group <strong>or</strong> its derivatives decreased during tree reaction. The<br />

concentration of flavonoids other than Pc increased <strong>or</strong> decreased depending on clone.<br />

Hydroxycinnamie<br />

Flavonoid Stilbene acid group <strong>or</strong> its<br />

group group derivatives<br />

(a) (a) (a)<br />

Figure 2.mVariations in the concentration of the phenolic compounds, between unwounded phloem and reactive phloem of<br />

different Scots pine clones, after inoculation with Ophiostoma brunneo-ciliatum. (a) = compound appearing in the<br />

reaction zone. One arrow means that the direction of variation was the same f<strong>or</strong> all clones; several arrows mean that<br />

the direction of variation depended on the clone. Txf = taxifolin; Txfgl = taxifolin glycosid; Pc = pinocembrin;<br />

Ps = pinosylvin; Psme = pinosylvin monomethylether; Pcae = p-coumaric acid ester; Acphe = acetophenone glycoside.<br />

These results confirm the preliminary study by Lieutier et al. (1991a) in Scots pine. Phloem phenolic response to<br />

inoculations differed greatly acc<strong>or</strong>ding to compound, which caused dramatic modifications in the relative composition of <strong>this</strong><br />

tissue. These results differ completely from the observations on terpenes in any conifer species (Russel and Berryman 1976;<br />

Raffa and Berryman 1982a,b; Del<strong>or</strong>me and Lieutier 1990; Lieutier et al. 1991b; Lfingstr6m et al. 1992), and makes certain<br />

phenols good candidates to play a role in arresting invaders and to be used as biochemical markers of tree resistance.<br />

Stilbenes and Pc, which were characteristic of phloem induced reactions, could play a particular role in reaction<br />

efficiency, and consequently in a tree's induced resistance to bark beetle and fungi attacks. However, similar results have<br />

been rep<strong>or</strong>ted in needles after pollution by ozone (Rosemann et al. 1991), as well as in the sapwood of various pine species in<br />

response to several kinds of aggression (Ref. in Kuc and Shain 1977, Hart and Shrimpton 1979, Shain 1979, Hart 1981,<br />

Kemp and Burden 1986). In addition, similar results have also been obtained with a phloem wound without fungus (Lieutier<br />

and Yart unpubl.). Lieutier et al. (1991a) also mentioned the non-specificity of the phloem phenolic response after various<br />

aggressions. Tree response to bark beetles and their associated fungi are thus m<strong>or</strong>e a response to wounding than a response to<br />

particular aggress<strong>or</strong>s themselves (Lieutier 1993). Mullick (1977) has suggested previously that a tree's response to aggres.<br />

sion is m<strong>or</strong>e concerned with tissue rest<strong>or</strong>ation than with defense.<br />

180


Considering biochemical pathways, it is interesting to note that the appearance of stilbenes and Pc in Scots pine<br />

c<strong>or</strong>responded to decreases in compounds of the hydroxycinnamic acid group, since <strong>this</strong> latter group is a precurs<strong>or</strong> of both<br />

stilbenes and flavonoids (Ribereau-Gayon 1968, Gotham 1989).<br />

Between Clone Comparisons<br />

Clonal variability was first analyzed by multivariate analysis, in <strong>or</strong>der to take into account all characterized com-<br />

pounds together. A canonical discriminant analysis was perf<strong>or</strong>med by taking reaction zones as Inain individuals and un-<br />

wounded phloem as complementary individuals. This procedure was necessary as the purpose was mainly to compare clone<br />

:reaction zones with each other, and only to place unwounded phloem relatively to them.<br />

Axes 1 2 and 3 explained 57, 23, and 15% of the variation, respectively. In the first two axes (Fig. 3), froln un-<br />

'<br />

wounded phloem to the reaction zone, phenolic composition seemed to evolve towards the same point of the plane, whatever<br />

the clone. That was confirmed in all significant axes (three), by calculating mean distances between points and their<br />

barycenter. Spread of the points relating to reaction zones was less than spread of the points relating to unwounded phloeln,<br />

in each clone and f<strong>or</strong> all clones together (data not shown). Thus, each clone would react in a different way acc<strong>or</strong>ding to its<br />

constitutive composition, but always so that the composition of the reaction zone tended to be the same.<br />

The meaning of <strong>this</strong> phenolic modifications is not clear. It may be linked with protection against aggress<strong>or</strong>s <strong>or</strong> with<br />

tissue rest<strong>or</strong>ation. Nevertheless, it is w<strong>or</strong>thwhile to note that clone ...... c , whose reaction was the least efficient (reaction zone<br />

significantly longer than that of' all other clones), was the only one initially separated from the others by axis 2 (Fig. 3). If<br />

phenols effectively play a role in arresting beetles and their fungi, it could thus be possible to predict reaction efficiency of a<br />

clone on the basis of the whole phenolic composition of its constitutive phloem.<br />

8<br />

_'.%0<br />

eql<br />

-8<br />

.<br />

d<br />

""N&<br />

2 3,<br />

4 4<br />

/ 5<br />

b'b_za 6 "_so_ s<br />

I I I II<br />

-10 0 10<br />

Axis 1 (57%)<br />

Figure 3._Canonical discriminant analysis of phloem phenolic composition of different Scots pine clones, after inoculation<br />

with Ophiostoma brunneo-ciliatum. Reaction zones were considered as principal individuals and unwounded phloem<br />

as complementary individuals, a, b, c, d, e, f = unwounded phloem of the different clones (* represents the<br />

barycenters); 1, 2, 3, 4, 5, 6 = C<strong>or</strong>responding reaction zones (o represent the barycenters). Arrows indicate the<br />

directions of variation from unwounded phloems to reactions zones. Clones a, b, c, d, e and f were respectively<br />

represented by 3, 2, 1, 2, 3 and 3 trees.<br />

t81


Clonal variability was also analyzed compound by compound, f<strong>or</strong> the peaks characterized at least at the family level.<br />

To appreciate if certain compounds could be used as markers of efficiency of tree response, phenol concentration in each<br />

clone and f<strong>or</strong> each compound, was compared to reaction zone length. The best c<strong>or</strong>relation (r = 0.78) was f<strong>or</strong> taxifolin<br />

glycoside in unwounded phloem (data not shown). The lowest concentrations c<strong>or</strong>responded to the least extended reaction<br />

zones, that is to the most effective response, and inversely. Concentration of taxifolin glycoside could thus be a marker of<br />

tree's response efficiency. Presence of taxifolin glycoside in the phloem pri<strong>or</strong> to aggression could be fav<strong>or</strong>able to the fungus,<br />

<strong>or</strong> could directly impede the development of the tree response. In the two cases, the result would be an increase in the lag<br />

time necessary to stop the fungus, thus causing a long reaction zone.<br />

Kinetics of the Phenolic Response<br />

A first approach to the kinetics was perf<strong>or</strong>med with multivariate analysis, the same way as in the previous experiment<br />

(Fig. 4). Axes 1 and 2 explained 88 and 9 percent of the variation, respectively. The composition of all unwounded samples<br />

was close to that of reaction zones at day 3, which suggest that phenolic composition of unwounded phloem did not vary<br />

3<br />

1<br />

4 u<br />

,,_,,, ..t 4<br />

2<br />

3<br />

4<br />

333 5 S<br />

2434 5<br />

At ed s<br />

-3 d<br />

[<br />

-3 ' -i " i ' ; '.... i '<br />

Axis 1 (88%)<br />

Figure 4.--Canonical discriminant analysis of the phenolic composition of Scots pine phloem collected at different dates after<br />

inoculations with Ophiostoma brunneo-ciliatum, in three different trees. Reaction zones were considered as principal<br />

individuals and unwounded phloem as complementary individuals. 1, 2, 3, 4, 5 = Reaction zones collected respectively<br />

3, 7, 14, 30 and 60 days after inoculations. A, B, C, D, E = c<strong>or</strong>responding barycenters, d, e = unwounded<br />

phloem collected at days 30 and 60. Arrows indicate the directions of variation during reaction development.<br />

during the whole experiment. Although between tree variability existed, the phenolic composition of the reaction zones at<br />

day 60 was clearly separated from that of all other samples along axis 1. Reaction zones at days 3 were separated from other<br />

samples along axis 2. Kinetics of phenolic response thus proceeded in two main phases. Concomitantly, growth of the<br />

reaction zone also proceeded in two main phases and was approximately done at day 30, indicating that the fungus had been<br />

stopped bef<strong>or</strong>e <strong>this</strong> date (Fig. 5). One can thus hypothesize that the first phase, bef<strong>or</strong>e day 30, is related to resistance and to<br />

interactions between tree and fungus, while the second m<strong>or</strong>e probably c<strong>or</strong>responds to wound healing processes. Considenng<br />

that, it is interesting to note that wound periderm began to be visible at day 30.<br />

Analysis, compound by compound, revealed three main kinds of kinetics. Ps, Psme and Pc increased in two phases,<br />

first early until day 7, then late after day 30 (Fig. 6a and 6b). Some other compounds, such as acetophenone glycoside,<br />

regularly decreased <strong>or</strong> remained constant (Fig. 6c), while others, such as taxifolin, increased only late, after day 30, SOmetimes<br />

after a decrease at the very beginning of development of the reaction (Fig. 6d).<br />

182<br />

5<br />

, i: i'i:i_i_'i!ill


- =<br />

120<br />

N /'_"<br />

=40 1/<br />

_<br />

j Days<br />

0 _<br />

20 40 60<br />

0 ..___... tree A .-- m tree B ....... tree C<br />

Growth of the reaction zone<br />

Figure 5.--Mean reaction zone length at different dates after inoculation of three trees with Ophiostoma brunneo-ciliatum.<br />

Verticalbars represent standard deviations.<br />

.N<br />

k,<br />

_0<br />

0,5<br />

3'<br />

0,4 Pinocembrin t Pinosylvin monomcthylethcr<br />

-;--f_- o , J<br />

_ 0 _0 io 3'0 40 _0 60 0 _0 20 30 40 50 60<br />

° acctophcnone Taxifolin<br />

_ 15 glycoside . 4<br />

,o.., 1<br />

",L1s _l<br />

"''l "_I --__ i 2<br />

................ RZ ............................................ UP ] "_"" ....................... 17_--=='2"- Days<br />

0 ..................................................................<br />

" 0 " _i" _"_''_<br />

o io io 3b 4b ._'o go 0 fo 2'0 _0 4'0 s'o 6'o<br />

Tree A Tree B ................... Tree C<br />

UP unwounded phloem RZ reaction zone<br />

Figure 6._Variations of the mean concentration of various phenolic compounds in reacting phloem and in unwounded phloem<br />

of three trees, after inoculation with Ophiostoma brunneo-ciliatum. Vertical bars represent standard deviations.<br />

183


Under these conditions, acc<strong>or</strong>ding to the above hypothesis, Ps, Psme and Pc would probably be involved in tree<br />

resistance, which could be related to the tree's ability to rapidly synthesize these compounds. After an early increase, their<br />

concentration remains constant until the fungus was stopped, which could indicate that they were metabolized by the fungus<br />

because of their toxic effect, as suggested by Lyr (1962). Their increase after day 30 could thus be due to the absence of<br />

fungus activity. This increase also suggests that they are involved in wound healing processes, as could be taxifolin. In Sitka<br />

Spruce as well as in hardwood trees, accumulation of phenolic compounds takes place parallel to suberin biosynthesis, a<br />

component of the wound periderm (Biggs 1985, Woodward and Pearce 1988).<br />

CONCLUSIONS<br />

Tree response could be m<strong>or</strong>e a generalized wound response than a pure defensive response. Both stilbenes and<br />

flavonoids could be involved in tree resistance. Taxifolin glycoside is a potential marker of tree resistance, all the m<strong>or</strong>e<br />

interesting as it belongs to constitutive phloem, but contrary to pinocembrin and stilbenes, its high concentration would<br />

indicate a low level of resistance. Whole phenolic composition of unwounded phloem could also be a marker of resistance.<br />

However, several conditions need to be verified bef<strong>or</strong>e definitive conclusions are possible about their use as markers. Indeed,<br />

between clone differences could depend on tree physiological status <strong>or</strong> age, environmental conditions, locality <strong>or</strong> season. In<br />

addition, if genetic markers are looked f<strong>or</strong>, the c<strong>or</strong>responding compounds must be genetically dependent and the c<strong>or</strong>responding<br />

genes must always be expressed.<br />

SUMMARY<br />

Scots pine phloem was investigated f<strong>or</strong> phenolic markers of tree resistance to bark beetles. Our approach was based<br />

on the study of relations between the reaction efficiency and phenolic composition in six different clones, and on the study of<br />

the variations in phenolic composition during tree reaction development. Considerable modifications took place after<br />

aggression. Compounds lacking in unwounded phloem appeared in the reaction zone. During the development of the<br />

reaction zone, modifications of the phenolic composition clearly proceeded in two successive phases. Stilbenes and flavonoids<br />

seemed to be involved in the induced resistance. Taxifolin glycoside in unwounded phloem looked interesting as<br />

marker, as did the whole phenol composition of <strong>this</strong> tissue.<br />

ACKNOWLEDGMENTS<br />

The auth<strong>or</strong>s thank J. Garcia and P. Romary (INRA, Orleans, France) f<strong>or</strong> their technical help and Dr. D. Sauvard<br />

(INRA, Orldans) f<strong>or</strong> help with the statistics. They are grateful to Prof. H. Sandermann and Dr. W. Heller (G.S.F. Neuherberg,<br />

Deutschland) and to Dr. A. Scalbert (INRA Grignon, France) f<strong>or</strong> providing pure compounds. They also thank Drs B.<br />

Monties, A. Scalbert and M.T. Tollier (INRA Grignon, France)/'<strong>or</strong> their help in characterization. These studies were granted<br />

by the Conseil R6gional de la Rdgion Centre (France) and by the French Ministry f<strong>or</strong> <strong>Research</strong> and Technology.<br />

LITERATURE CITED<br />

ALCUBILLA, M. 1970. Extraction, chromatographic separation and isolation of fungistatic substances from the inner bark<br />

of N<strong>or</strong>way Spruce. Z. Pflanz. Bodenk. 127: 64-74.<br />

BERRYMAN, A.A. 1972. Resistance of conifers to invasion by bark beetle fungus associations. BioScience 22:598-60 1.<br />

BERRYMAN, A.A. 1978. A synoptical model of the lodgepole pine/mountain pine beetle interaction and its implication in<br />

f<strong>or</strong>est management, p. 98-105. In Berryman, A.A., Amman, G.D., Stark, R.W., and Kibbee, D.L., eds. The<strong>or</strong>y and<br />

Practice of Mountain Pine Beetle Management in Lodgepole Pine F<strong>or</strong>ests. Coll. F<strong>or</strong>. Res., Univ. Idaho, Moscow.<br />

BERRYMAN, A.A. 1982. Biological control, thresholds and pest outbreaks. Environ. Entomol. 11: 544-549.<br />

184


BIGGS, A.R. 1985. Suberized boundary zones and the chronology of wound response in tree bark. Phytopathol. 75:1191-<br />

1195.<br />

CHRISTIANSEN, E, and HORNTVEDT, R. 1983. Combined Ips/Ceratoo'stis attack on N<strong>or</strong>way spruce, and defensive<br />

mechanisms of the trees. Z. Angew. Entomol. 96:110-118.<br />

CHRISTIANSEN, E., WARING, R.H., and BERRYMAN, A.A. 1987. Resistance of conifers to bark beetle attack: Search-<br />

ing f<strong>or</strong> general relationships. F<strong>or</strong>. Ecol. Manage. 22: 89-106.<br />

DELORME, L. and LIEUTIER, F. 1990. Monoterpene composition of the pref<strong>or</strong>med and induced resins of Scots pine, and<br />

their effect on bark beetles and associated fungi. Eur. J. F<strong>or</strong>. Pathol. 20:304-316.<br />

GORHAM, J. 1989. Stilbenes and Phenanthrenes, p. 159-196. In Dey, M. and Harb<strong>or</strong>ne, J.B., eds. Methods in Plant<br />

Biochemistry. Vol. 1, Plants phenolics, acad. Press, MY.<br />

HART, J.H. 1981. Role of phytostilbenes in decay and desease resistance. Ann. Rev. Phytopathol. 19: 437-458.<br />

HART, J.H. and SHRIMPTON, D.M. 1979. Role of stilbenes in resistance of wood to decaY. Phytopathol. 69: 1138-1143.<br />

KEMR M.S. and BURDEN, R.S 1986. Phytoalexins and stress metabolites in the sapwood of trees. Dhytochemistry 25"<br />

1261-1269.<br />

KUC, J. and SHAIN,L. 1977. antifungal compounds associated with disease resistance in plants, p. 497-535. In Siegel,<br />

M.R. and Sisler, H.D., eds. antifungal Compounds. Vol. 2, Marcel Dekker, MY.<br />

LANGSTROM, B., HELLQVIST, C., ERICSSON, A., and GREF, R. 1992. Induced defence reaction in Scots pine follow-<br />

ing stem attacks by Tomicus piniperda. Ecography 15: 318-327.<br />

LIEUTIER, F. 1993. Induced defense reactions of conifers to bark beetles and their associated Ophiostoma, p. 206-215. In<br />

Wingfield, M.J., Seifert, K.A., and Webber, J.F., eds. Ceratocystis and Ophiostoma: Taxonomy, Biology and Pathogenicity.<br />

Am. Phytopathol. Soc., St Paul.<br />

LIEUTIER, F., YART, A., GARCIA, J., POUPINEL, B. and LEVIEUX, J. 1988 Do fungi influence the establishment of<br />

,<br />

bark beetles on Scots pine? p. 317-330. In Mattson, W.J., Levieux, J., and Bernard-dagan, C., eds. Mechanisms of<br />

Woody Plant Defenses Against insects, Search f<strong>or</strong> Pattern. Springer, MY.<br />

LIEUTIER, F., YART, A., JAY-ALLEMAND, C., and DELORME, L. 1991a. Preliminary investigations on phenolics as a<br />

response of Scots pine phloem to attacks by bark beetles and their associated fungi. Eur. J. f<strong>or</strong>. Pathol. 21: 354-364.<br />

LIEUTIER, F., BERRYMAN, A.A., and MILLSTEIN, J.A. 1991b. Preliminary study of the monoterpene response of three<br />

pines to Ophiosloma clavigerum (ascomycetes: Ophiostomatales)and two chemical elicit<strong>or</strong>s. Ann. Sc. F<strong>or</strong>. 48: 377-<br />

388.<br />

LIEUTIER, F., GARCIA, J., ROMARY, R, YART, A., JACTEL, H., and SAUVARD, D. 1993. Inter-tree variability in the<br />

induced defense reaction of Scots pine to single inoculations by Ophiostoma brunneo-ciliaturn, a bark beetle associated<br />

fungus. F<strong>or</strong>. Ecol. Manage. 59: 257-270.<br />

LIEUTIER, F., PICRON, V., BRIGNOLAS, F., YART, A., SAUVARD, D., BASTIEN, C., and JAY-ALLEMAND, C. 1995.<br />

Changes in soluble phenolic metabolites of Scots pine phloem in response to challenge with Ophkvstoma brunneociliatum,<br />

a bark beetle associated fungus. Eur. J. f<strong>or</strong> Pathol. in press.<br />

LYR, H. 1962. Detoxification of heartwood toxins and chl<strong>or</strong>ophenols by higher fungi. Nature 194: 289-290•<br />

MULLICK, D.B. 1977. The non-specific nature of defence in bark and wood during wounding insect and pathogen attack,<br />

Recent Adv. Phytochem., Vol. 2, Plenum, New Y<strong>or</strong>k: 395-441.<br />

185


RAFFA, K.E and BERRYMAN, A.A. 1982a. Accumulation of monoterpenes and associated volatiles following inoculation<br />

of grand fir with a fungus transmitted by the fir engraver, Scolytus ventralis (Coleoptera: Scolytidae). Can. EntomoI,<br />

114:797-810.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1982b. Physiological differences between lodgepole pines resistant and susceptible to<br />

the mountain pine beetle and associated micro<strong>or</strong>ganisms. Environ. Entomol. 2: 486-492.<br />

REID, R.W., WHITNEY, H.S., and WATSON, J.A. 1967. Reactions of lodgepole pine to attack by Dendroctonus<br />

ponderosae Hopk. and blue stain fungi. Can. J. Bot. 45:1 1115-1126.<br />

RIBEREAU-GAYON, R 1968. Les compos6s ph6noliques des v6g6taux. Dunod ed., Paris. 253 p.<br />

ROSEMANN, D., HELLER, W., and SANDERMANN, H. 1991. Biochemical plant response to ozone. II. Induction of<br />

stilbene biosynthesis in Scots pine (Pinus sylvestris L.) seedlings. Plant Physiol. 97: 1280-1286.<br />

RUSSEL, C.E. and BERRYMAN, A.A. 1976. Host resistance to the fir engraver beetle: 1 - Monoterpene composition of<br />

Abies grandis pitch blisters and fungus-infected wounds. Can. J. Bot. 54: 14-18.<br />

SHAIN, L. 1979. Dynamic responses of differentiated sapwood to injury and infection. Phytopathol. 69:1143-1147.<br />

SHRIMPTON, D.M. 1973. Extractives associated with the wound response of lodgepole pine attacked by the mountain pin<br />

beetle and associated micro<strong>or</strong>ganisms. Can. J. Bot. 51: 527-534.<br />

WOODWARD, S. and PEARCE, R.B. 1988. Wound-associated response in Sitka Spruce root bark challenged with<br />

Phaeolus schweinitzii. Physiol. Molecular Plant Pathol. 33:151-162.<br />

186


DIFFERENTIAL SUSCEPTIBILITY OF WHITE FIR PROVENANCES<br />

TO THE FIR ENGRAVER AND ITS FUNGAL SYMBIONT<br />

IN NORTHERN CAHFORNIA<br />

G. T. FERRELL _and W. J. OTROSINA 2<br />

_<strong>USDA</strong> F<strong>or</strong>est Service, Pacific Southwest <strong>Research</strong> <strong>Station</strong>, 2600 Washington Avenue, Redding, Calif<strong>or</strong>nia 96001, USA<br />

2<strong>USDA</strong> F<strong>or</strong>est Service, Southeastern F<strong>or</strong>est Experiment <strong>Station</strong>, 320 Green Street, Athens, Ge<strong>or</strong>gia 30602, USA<br />

INTRODUCTION<br />

The fir engraver, Scolytus ventralis LeC., attacks white fir, Abies concol<strong>or</strong> (G<strong>or</strong>d. and Glend.) Lindl., and other true<br />

firs, Abies spp., in western N<strong>or</strong>th America. The biology, attack behavi<strong>or</strong>, and ecology of <strong>this</strong> bark beetle were recently<br />

summarized by Berryman and Ferrell (1988). During the summer flight season, the attacking beetles b<strong>or</strong>e into the cambial<br />

zone of fir boles, introducing a pathogenic brown-staining fungus, Trichosp<strong>or</strong>ium symbioticum Wright. Resistant firs react to<br />

the invasion by f<strong>or</strong>ming a resinot_s necrotic wound in the phloem and outer sapwood, which contains the spread of the fungus<br />

and repels <strong>or</strong> kills the beetles. In such interactions, the tree usually survives. This reaction is less intense <strong>or</strong> absent in<br />

susceptible firs, resulting in reproductive success of the beetles, and severe damage <strong>or</strong> even death of the fir. Sp<strong>or</strong>adic<br />

outbreaks of the fir engraver, associated primarily with droughts, have caused widespread m<strong>or</strong>tality of true firs in nearly<br />

every decade of <strong>this</strong> century in western N<strong>or</strong>th America.<br />

In the drought years of 1987-88, four 26-year-old white fir provenance test plantations at Camino, Calif<strong>or</strong>nia, located<br />

at 1028 m elevation in the central Sierra Nevada, sustained considerable levels of m<strong>or</strong>tality. Subc<strong>or</strong>tical examination of a<br />

subsample of the dead firs indicated that all had reproductively successful gallery systems of the fir engraver. Drought<br />

continued until winter 1992-1993. Surveys beginning in fall 1988 revealed pronounced differences in <strong>this</strong> m<strong>or</strong>tality, among<br />

both plantations and provenances, and also within plantations.<br />

This paper describes the observed patterns of susceptibility and resistance in relation to known patterns of geographic<br />

variation in white fir in western N<strong>or</strong>th America and discusses studies underway to understand the mechanisms responsible f<strong>or</strong><br />

these patterns.<br />

METHODS<br />

The adjacent provenance test plantations represent two studies. The "geographic range" plantations (1 and 3) consist<br />

of 39 provenances from throughout most of the western p<strong>or</strong>tion of white fir's natural geographic range in western N<strong>or</strong>th<br />

America. Each provenance was <strong>or</strong>iginally represented by three replications, each of three seedlings. One replication was<br />

removed by thinning in 1970. The "elevational transect" plantations (2 and 4) contain only four provenances representing a<br />

west-east elevational transect across the Sierra Nevada at the latitude of Camino. Each of these provenances is represented<br />

by 10 half-sib families, each <strong>or</strong>iginally with nine trees, since thinned to six. In all four plantations, provenances were planted<br />

in an interlocked randomized non-contiguous plot layout designed to minimize effects of microsite variation (Libby and<br />

Cockerham 1980).<br />

Beginning in August 1988, trees in these plantations were surveyed annually in summer <strong>or</strong> early fall f<strong>or</strong> m<strong>or</strong>tality and<br />

susceptibility to fir engravers. Surveys were conducted after trees attacked and killed by fir engravers the previous year had faded<br />

crowns, but bef<strong>or</strong>e trees killed by current attack had faded crowns. During these surveys, trees topkilled by the previous year's<br />

attack were also noted.<br />

Mattson, W.J., Niemel/i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle.<br />

<strong>USDA</strong> F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.


Patterns of tree resistance were surveyed in December 1989; numbers of pitch streamers indicating unsuccessful fir<br />

engraver attacks were classified as 0, < 5, 5-10, and > 10 on stems of all live trees.<br />

Distributions of trees classified by status (live, dead) and number of pitch streamers were compared among provenances<br />

by Chi-square Contingency Test (SAS Institute Inc. 1988). Data from the two replicates of each type of plantation<br />

(geographic range, elevational transect) were similar and were pooled f<strong>or</strong> analysis. In the geographic range plantations,<br />

provenances were combined acc<strong>or</strong>ding to Hamrick and Libby (1972) as modified by Libby et al. (1980). The latter auth<strong>or</strong>s<br />

divided white fir in the western part of its range into five maj<strong>or</strong> geographic groups based primarily on needle m<strong>or</strong>phology:<br />

(1)"N<strong>or</strong>thern"--central Oregon and n<strong>or</strong>thwestern Calif<strong>or</strong>nia; (2) "Central"-- south-central Oregon, central and n<strong>or</strong>theastern<br />

Calif<strong>or</strong>nia; (3) "Southern Calif<strong>or</strong>nia"; (4) "Interi<strong>or</strong> South"m Arizona; and (5) "Interi<strong>or</strong> N<strong>or</strong>th"-- eastern Nevada and western<br />

Utah.<br />

RESULTS<br />

Of the 1,127 firs alive in all four plantations in 1987, 393 (35%) were killed by the fir engraver by 1993. Most (253)<br />

of the m<strong>or</strong>tality resulted from the attack season of 1987, with an additional 121 resulting from the 1988 attack season.<br />

Relatively few (29) were killed by attack in 1989-1993. Only a few (10) trees were topkilled, two of which were killed in<br />

subsequent years. Firs on the outside edges of the plantations were infrequently killed (4/97 <strong>or</strong> 4%).<br />

Through 1993, cumulative m<strong>or</strong>tality in the elevational transect plantations (256 of 578 total trees, <strong>or</strong> 44%) was almost<br />

twice that in the geographic range plantations (137/549, <strong>or</strong> 25%). The associated Chi-square (adjusted f<strong>or</strong> continuity) was<br />

45.505, p < 0.0001 at 1 dr.<br />

In the geographic range plantations, m<strong>or</strong>tality differed among maj<strong>or</strong> geographic groups of Hamrick and Libby (1972)<br />

and Libby et al. (1980). In both plantations, provenances of the N<strong>or</strong>thern and Central groups had sustained the highest<br />

m<strong>or</strong>tality, while very few of the firs of the Interi<strong>or</strong> South and Interi<strong>or</strong> N<strong>or</strong>th groups were killed (Table 1). Low levels (1 1<br />

percent) of m<strong>or</strong>tality occurred in southern Calif<strong>or</strong>nia firs. The Chi-square associated with among-group variation was<br />

40.631, p < 0.0001 at 4 df. Among individual provenances, the one from nearest Camino (AK; Omo Ranch; 32 km distant)<br />

had the highest m<strong>or</strong>tality in both plantations.<br />

Table 1.--M<strong>or</strong>tality of white fir provenances caused by fir engravers in the<br />

geographic range plantations t at Camino, CA I987-1993<br />

Group 2 Plantation 1 Plantation 3<br />

Total Dead Pctdead Total Dead Pctdead<br />

N<strong>or</strong>thern 55 12 22 84 33 39<br />

Central 106 22 21 124 55 44<br />

Southern Calif. 40 3 8 47 7 15<br />

Interi<strong>or</strong> South 28 0 0 31 4 13<br />

Interi<strong>or</strong> N<strong>or</strong>th !4 0 0 20 1 5<br />

1Each plantation contained 39 provenances from the western p<strong>or</strong>tion of white fir's<br />

geographic range.<br />

ZProvenances combined by maj<strong>or</strong> m<strong>or</strong>phological groups of Hamrick and Libby<br />

(1972).<br />

In the elevational transect plantations, m<strong>or</strong>tality in the two lower- elevation westside provenances nearest Camino<br />

(AK, AL) averaged m<strong>or</strong>e than one-third greater than that in the two upper-elevation eastside provenances (AM, AN), which<br />

are only about 50 km m<strong>or</strong>e distant (Table 2). The among-provenance Chi-square was 24.757, p < 0.0001 at 3 df.<br />

Analysis of the distribution of pitch streamers on boles of surviving firs revealed that in the elevational transect<br />

plantations, virtually all surviving firs had been attacked, and many heavily so (>10 streamers). In the geographic range<br />

188


Table 2.--M<strong>or</strong>tality of white fir provenances caused by fir engravers in the<br />

elevational transect plantations _at Camino, CA 1987-1993<br />

Provenance Plantation 2 Plantation 4<br />

Total Dead Pct dead Total Dead Pctdead<br />

AK 76 37 49 95 59 62<br />

AL 72 30 42 72 42 58<br />

AN 65 19 29 63 27 43<br />

AM 61 9 15 74 33 45<br />

_West-east transect across Sierra Nevada at latitude of Camino. AK, AL are<br />

lower elevation westside; AN, AM are upper elevation eastside.<br />

plantations, however, m<strong>or</strong>e than 38% of Interi<strong>or</strong> South and Interi<strong>or</strong> N<strong>or</strong>th firs had no pitch streamers, and fewer than 11%<br />

had m<strong>or</strong>e than l0 pitch streamers (Table 3). In contrast, among N<strong>or</strong>thern and Central firs, fewer than 6% had no pitch<br />

streamers while m<strong>or</strong>e than 24% had over 10. Distribution of pitch streamers on Southern Calif<strong>or</strong>nia firs was intermediate<br />

between these two patterns. The among-group Chi-square was 128.424, p < 0.0001 at 12 df. The Central provenances from<br />

the Sierra Nevada where Camino is located had the lowest percentage of firs with no pitch streamers (ca. 2%) and the highest<br />

percentage with m<strong>or</strong>e than 10 streamers (ca. 41%).<br />

Table 3.--Incidence of pitch streamers caused by fir engravers on boles of white fir<br />

provenances in the geographic range plantations, _Camino, 1989<br />

Group 2 Number of pitch streamers<br />

0 10<br />

N<strong>or</strong>thern 5 34 31 23<br />

Central 3 22 61 60<br />

Southern Calif. 11 16 33 16<br />

Interi<strong>or</strong> South 27 14 8 6<br />

Interi<strong>or</strong> N<strong>or</strong>th 13 12 8 1<br />

JPlantations 1 and 3, combined.<br />

2provenances combined by maj<strong>or</strong> m<strong>or</strong>phological groups of Hamrick and Libby (1972).<br />

DISCUSSION<br />

Results indicated that N<strong>or</strong>thern (Oregon) and Central (N<strong>or</strong>thern Calif<strong>or</strong>nia) provenances were susceptible to the fir<br />

engraver and its fungal symbiont in the Camino plantations, while Interi<strong>or</strong> South (Arizona), and Interi<strong>or</strong> N<strong>or</strong>th (Nevada,<br />

Utah) provenances were virtually nonsusceptible. Southern Calif<strong>or</strong>nia provenances demonstrated low susceptibility. These<br />

patterns agree closely with known geographic patterns of m<strong>or</strong>phological and chemical (c<strong>or</strong>tical monoterpenes) variation in<br />

white fir over its natural range in western N<strong>or</strong>th America (Hamrick and Libby 1972, Zavarin et al. 1975). Susceptible<br />

provenances were those characterized as green-foliaged, with needle m<strong>or</strong>phology suggesting only moderate drought resistance,<br />

and low in camphene and 3-carene, consisting of Calif<strong>or</strong>nia white fir (var. lowiana (G<strong>or</strong>d.) Lemm.) from n<strong>or</strong>thern<br />

Calif<strong>or</strong>nia and intermediates with grand fir, A. grandis (Doug.), from south-central Oregon. Provenances evidencing virtually<br />

no susceptibility were blue-green foliaged with needle m<strong>or</strong>phology suggesting higher drought resistance, relatively high<br />

in camphene and 3-carene, and of the Rocky Mountain variety (vat. concol<strong>or</strong> (G<strong>or</strong>d. & Glend.) Lindl.). Southern Calif<strong>or</strong>nia<br />

provenances evidencing low susceptibility were characterized as blue-green foliaged, relatively high in 3-carene but nearly<br />

lacking in camphene and thus intermediate between the N<strong>or</strong>thern and Interi<strong>or</strong> South groups.<br />

189


In all plantations, susceptibility was always highest in provenances nearest Camino. This was evident not only on a<br />

large scale in the geographic range plantations where Central (N<strong>or</strong>thern Calif<strong>or</strong>nia) provenances were the most susceptible,<br />

but also on a fine scale in the elevational transect plantations where provenances 40 km distant from Camino were m<strong>or</strong>e<br />

susceptible than those from 80 kin distant. In loblolly pine plantations in South Carolina, Powers et al. (1992) also found that<br />

local provenances were m<strong>or</strong>e susceptible to southern pine beetle, Dendroctonusfrontalis, than distant provenances.<br />

Mechanisms responsible f<strong>or</strong> the observed differential susceptibility among provenances are under investigation.<br />

Their basis, however, is undoubtedly primarily genetic as the plantations are all adjacent with provenances planted intermixed<br />

to minimize chances that microsite would differentially affect provenances.<br />

The very low susceptibility of trees on the outside edges of the plantations compared with those in the interi<strong>or</strong> was<br />

probably primarily determined by stand fact<strong>or</strong>s such as differences in soil moisture <strong>or</strong> beetle pheromone dispersal rather than<br />

tree genetics because susceptibility occurred without regard to provenance. Mechanisms underlying <strong>this</strong> difference are under<br />

investigation.<br />

We have two studies underway testing the hypothesis that observed differential susceptibility is attributable to<br />

differential moisture stress among provenances, and between firs on plantation edge versus interi<strong>or</strong>. White firs with highly<br />

negative water potentials (pre-dawn in August during summer dry, and fir engraver flight, seasons) are known to be susceptible<br />

to the fir engraver (Ferrell 1978). Results to date failed to find appreciable differences among provenances in 1989,<br />

1990, <strong>or</strong> 1993. "Edge" firs did average higher moisture stress than "interi<strong>or</strong>" firs in 1992 and 1993. Both studies will<br />

continue f<strong>or</strong> one m<strong>or</strong>e post-drought year. Results are necessarily conditioned by the fact that moisture stress can be studied<br />

only in surviving firs. Thus, parallel studies are underway to compare radial growth patterns in surviving versus killed firs,<br />

under the hypothsis that differential moisture stress should be evident in radial growth pattterns.<br />

Differences in fir engraver attack preference and success are being investigated by reciprocally caging "local" and<br />

"exotic" populations of beetles with bolts cut from "local" and "exotic" provenances of white fir. These tests are designed to<br />

reduce <strong>or</strong> eliminate differential resistance caused primarily by differences in environmental fact<strong>or</strong>s such as moisture stress<br />

and to isolate f<strong>or</strong> analysis differential resistance caused by fact<strong>or</strong>s that are primarily genetically determined such as constitutive<br />

bark chemicals. Preliminary results indicate that beetles from the vicinity of Camino initiate far fewer attacks in Arizona<br />

bolts than in "local" bolts. No such preference was evident in tests with Nevada beetles caged with bolts of firs from Nevada<br />

and the vicinity of Camino. These tests are being repeated with beetles and bolts of other provenances. Parallel studies are<br />

also under way using various geographic isolates of the Trichosp<strong>or</strong>ium fungus inoculated into stems of surviving firs in the<br />

plantations. Among-provenance variations in inoculation wound size and monoterpene composition are being studied. Thus<br />

far consistent differences in virulence have been tbund between isolates regardless of provenance of the fungus <strong>or</strong> the fir.<br />

Studies continue to expl<strong>or</strong>e seasonal and yearly variations and to analyze monoterpene composition.<br />

Results indicate that local populations of host conifers can be m<strong>or</strong>e susceptible to local populations of bark beetles<br />

and their fungal symbionts while exotic host populations can be less susceptible. If generally true, <strong>this</strong> phenomenon may<br />

have to be taken into account in tree improvement programs, most of which prefer to utilize local genetic material as planting<br />

stock because <strong>this</strong> material is preadapted to the growing site. If results at Camino are any indication, however, exotic seed<br />

sources may be less susceptible to local bark beetle populations provided that these trees are adequately adapted to the<br />

physiographic fact<strong>or</strong>s of the growing site. At Camino, many of the nonsusceptible exotic provenances had grown fully as tall<br />

as the best-growing local provenances, indicating that they were thus far as welt adapted to the growing site.<br />

Another imp<strong>or</strong>tant implication of the Camino results is that maintenance of genetic diversity in conifer populations<br />

may be an imp<strong>or</strong>tant safeguard in protecting conifer stands against bark beetles. The much greater m<strong>or</strong>tality experienced in<br />

the elevational transect plantations containing only four provenances may be an expression of <strong>this</strong>, although it was evidently<br />

also influenced by the local <strong>or</strong>igin of all four of these provenances.<br />

t90


Results from the geographic range plantations suggest that planting well-adapted exotic and local provenances in<br />

mixture may be a useful strategy f<strong>or</strong> avoiding problems from bark beetles and their fungal symbionts. Any use of exotic<br />

provenances should probably be limited to high-value trees, particularly those grown on stressful sites, because m<strong>or</strong>e widespread<br />

planting could lead to loss of resistance through local pest populations becoming adapted to them.<br />

SUMMARY<br />

During a drought-associated fir engraver outbreak in Calif<strong>or</strong>nia, local white fir provenances were m<strong>or</strong>e susceptible<br />

than exotic provenances, and, doubtless partly in consequence, less genetically diverse plantations were m<strong>or</strong>e susceptible<br />

than plantations with greater genetic diversity.<br />

Mechanisms underlying the observed differential susceptibility remain unknown but are the subject of continuing<br />

investigations.<br />

Results suggest that planting a mixture of well-adapted exotic, as well as local, provenances f<strong>or</strong> maintenance of high<br />

genetic diversity may be an imp<strong>or</strong>tant strategy f<strong>or</strong> protecting conifer hosts against bark beetles and their symbiotic fungi.<br />

ACKNOWLEDGMENTS<br />

We thank Dr. Tom Conkle and staff of the Pacific Southwest <strong>Research</strong> <strong>Station</strong>'s Institute of F<strong>or</strong>est Genetics and Prof.<br />

Bill Libby of the cooperating University of Calif<strong>or</strong>nia College of Environmental Science, Berkeley, f<strong>or</strong> access to plantations<br />

and f<strong>or</strong> encouragement.<br />

LITERATURE CITED<br />

BERRYMAN, A.A. and FERRELL, G.T. 1988. The fir engraver in western states, p. 555-577. In Berryman, A.A., ed.<br />

Dynamics of F<strong>or</strong>est Insect Populations. Plenum, New Y<strong>or</strong>k.<br />

FERRELL, G.T. 1978. Moisture stress threshold of susceptibility to fir engraver beetles in pole-size white firs. F<strong>or</strong>. Sci. 24:<br />

85-92.<br />

HAMRICK, J.L. and LIBBY, W.J. 1972. Variation and selection in western montane species. I. White fir. Silvae Genetica<br />

21" 29-35.<br />

LIBBY, W.J. and COCKERHAM, C.C. 1980. Random non-contiguous plots in interlocking field layouts. Silvae Genetica<br />

29: 183-190.<br />

LIBBY, W.J., ISIK, KANI, and KING, James P. 1980. Variation in flushing time among white fir population samples.<br />

Annales F<strong>or</strong>estales, Jugoslavenska Akademija Znanosti 8: 123-138.<br />

POWERS, H.R., JR., BELANGER, R.P., PEPPER, W.D., and HASTINGS, F.L. 1992. Loblolly pine seed sources differ in<br />

susceptibility to the southern pine beetle in South Carolina. South. J. Appl. F<strong>or</strong>. 16:169-174.<br />

SAS INSTITUTE INC. 1988. SAS/STAT User's Guide, Release 6.03 Edition. SAS Institute Inc. Cary, N<strong>or</strong>th Carolina.<br />

ZAVARIN, E., SNAJBERK, K., and FISHER, J. 1975. Geographic variability of monoterpenes from the c<strong>or</strong>tex of Abies<br />

concol<strong>or</strong>. Biochem. System. and Ecol. 3: 191-203.<br />

191


MILD DROUGHT ENHANCES THE RESISTANCE OF N©RWAY SPRUCE TO<br />

A BARK BEETLE-TRANSMITTED BLUE-STAIN FUNGUS<br />

ERIK CHRISTIANSEN and ANNIE MARIE GLOSLI<br />

N<strong>or</strong>wegian F<strong>or</strong>est <strong>Research</strong> Institute, N-1432 A.s, N<strong>or</strong>way<br />

INTRODUCTION<br />

Outbreaks of the spruce bark beetle, Ips typographus (L.), one of the most serious pests of the Eurasian spruce<br />

f<strong>or</strong>ests, are frequently triggered by large-scale windfelling. Although c<strong>or</strong>relative evidence indicates that long-lasting drought<br />

incites and aggravates such epidemics (cf. Christiansen and Bakke 1988), very little experimental w<strong>or</strong>k has been carried out<br />

to test <strong>this</strong> assumption.<br />

Ideally, the defensive capacity of the trees should be tested using prescribed numbers of bark beetle attacks. In Pinus<br />

cont<strong>or</strong>ta <strong>this</strong> has been done successfully by screening the branch-free part of the stem, and inducing a defined number of<br />

Dendroctonus ponderosae attacks under the screen (Raffa and Berryman 1983). Without screening, additional attacks by<br />

"wild" beetles in the neighbourhood will occur if the beetles possess aggregation pheromones.<br />

Unlike mountain pine beetle attacks on lodgepole pine, I. typographus attacks often extend far into the living crown<br />

of N<strong>or</strong>way spruce, Picea abies, making screening impractical. F<strong>or</strong> <strong>this</strong> reason we chose to assay host resistance using<br />

prescribed loads of artificial mass-inoculation with the blue-stain fungus, Ophiostoma polonicum Siem., a close associate of<br />

the spruce bark beetle.<br />

Sp<strong>or</strong>es of O. polonicum are transmitted both externally and internally by L typographus (Furniss et al. 1990), and the<br />

fungus is consistently isolated from the advancing front of blue-stain in successfully attacked trees (Solheim 1986, 11992). O.<br />

polonicum can kill healthy N<strong>or</strong>way spruce when artificially inoculated under the bark (H<strong>or</strong>ntvedt et al. 1983, Christiansen<br />

1985a, Solheim 1988). Similarly, other spruce species and even Pseudotsuga menziesii may succumb to mass-inoculation<br />

with <strong>this</strong> fungus (Christiansen and Solheim 1990). Parallel to the "threshold of successful attack" f<strong>or</strong> bark beetles<br />

(Thalenh<strong>or</strong>st 1958), a "threshold of successful infection" exists f<strong>or</strong> mass-inoculation with <strong>this</strong> fungus (Christiansen 1985b).<br />

Drought is generally accompanied by hot weather, and the two fact<strong>or</strong>s are not easily separated when one attempts to<br />

find the causes f<strong>or</strong> large-scale infestations. We hypothesize that (1) elevated temperatures act directly on the various life<br />

stages of the beetles to favour population build-up, and/<strong>or</strong> that (2) drought affects the physiology of the trees, making them<br />

m<strong>or</strong>e susceptible to attack <strong>or</strong> m<strong>or</strong>e suitable as food. Here, we rep<strong>or</strong>t an experiment designed to shed light on aspects of<br />

Hypothesis 2: N<strong>or</strong>way spruce trees were artificially drought stressed, and their defensive capability was assayed and compared<br />

with unstressed, control trees.<br />

In <strong>or</strong>dinary f<strong>or</strong>est stands, trees will vary considerably in their susceptibility to beetle/fungus attack. This necessitates<br />

large numbers of experimental trees to obtain statistically significant results. In the present experiment, however, we utilized<br />

clones of spruce trees and thus permined w<strong>or</strong>k with relatively few trees.<br />

METHODS<br />

Experimental Trees and Their Treatment<br />

The N<strong>or</strong>way spruce clones used in <strong>this</strong> study grew in a multi-clonal stand at Hogsmark Experimental Farm in ,As,<br />

Akershus, N<strong>or</strong>way, and had been produced as follows: Seeds from selected trees were sown in 1951 and cuttings from the<br />

Mattson, W.J., Nieme1_i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle.<br />

<strong>USDA</strong> F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St:.Paul, MN 55108.<br />

192


esulting seedlings were then rooted. In spring, 1970, rooted ramets of each clone were planted using 2x2 m spacing. The<br />

planting was done in a west-facing, gently sloping agricultural field where underground drain pipes had been installed and the<br />

clayey soil tilled bef<strong>or</strong>e planting. The homogeneous soil and the exact planting array reduced phenotypical variation in the<br />

stand to a minimum.<br />

From each of two clones, #194 and #582, twelve ramets were selected. Trees of Clone 194 had been used in an<br />

earlier mass-inoculation experiment with O. polonicum, and had proved relatively resistant to the infection (Christiansen,<br />

unpubl.). Nothing was known about the susceptibility of Clone 582 as to its susceptibility.<br />

The continuous rows of clones followed the slope of the field. The row of trees between Clone 194 and Clone 582<br />

had been cut in early October 1992, and so had the two rows on their other side. Thus both rows had a distance of 4 m to<br />

their nearest neighbours. Trees at both ends of the two rows were discarded as phenotypically aberrant, having much larger<br />

crowns than trees inside the stand.<br />

In late October, 1992, all branches below ca. 2.5 m were removed from all 24 trees. These branches were all dead<br />

due to shading. The ground on both sides of the upper six trees in each row was then covered with plastic sheets to prevent<br />

autumn and winter precipitation from percolating into the ground. Instead, <strong>this</strong> water was led down to the lower six trees in<br />

the two rows. On 20 April t993 the plastic ground cover was removed. Ropes were strung at 2 m above ground between the<br />

upper six trees of the rows, and non-transparent plastic tarpaulins were suspended from these ropes down to drains fastened<br />

to the stumps of the tree rows that had been removed. The drains collected the rain falling through the canopy, and channelled<br />

it down to the lower six trees in the rows.<br />

This way the six upper trees in a row were deprived of most of the precipitation falling from October till the end of<br />

the experiment. The lower six trees of the two rows served as unstressed controls. Because the early summer of 1993 was<br />

very dry, these control trees were given an extra 47 mm of water, applied under the canopy by means of a garden sprinkler.<br />

Monit<strong>or</strong>ing of Drought Stress<br />

Starting on 12 May 1993, drought stress was monit<strong>or</strong>ed by repeated measurements of pre-dawn xylem water potentials<br />

(hereafter termed "WP"), using a pressure chamber. Freshly excised shoots taken 4-5 m above ground were used. To<br />

detect a possible natural WP gradient along the slope, trees at the upper and lower ends of the rows were measured. By mid-<br />

June WPs of the experimental trees were clearly different from those of the controls, and after a monit<strong>or</strong>ing of all 24 trees on<br />

23 June we decided that they had reached a satisfact<strong>or</strong>y level of stress f<strong>or</strong> inoculation with the fungus.<br />

Fungal Inoculation<br />

Pri<strong>or</strong> to inoculation a 80 cm long template was attached around the stern between 1.2 and 2.0 m above ground. In the<br />

template evenly spaced holes had been punched at a density of 4 per dm 2,and the points of inoculation were marked through<br />

these holes. The choice of <strong>this</strong> particular dose was based on earlier experience with the susceptibility to infection by trees of<br />

Clone 194 and of two other clones growing in the same stand. Inoculation occurred on 25 June. Bark plugs were removed<br />

with a 5 mm c<strong>or</strong>k b<strong>or</strong>er. The inoculum, actively growing O. polonicum mycelium on malt agar, was placed in the cambiumholes,<br />

whereupon the bark plug was returned to its <strong>or</strong>iginal position (cf. H<strong>or</strong>ntvedt et al. 1983).<br />

Further Field and Lab<strong>or</strong>at<strong>or</strong>y Procedures<br />

Around the time of inoculation, exudation of constitutive resin was estimated twice (cf. Christiansen and H<strong>or</strong>ntvedt<br />

1983): on each tree 10 thin plastic tubes were inserted into holes cut with a c<strong>or</strong>k b<strong>or</strong>er about 2.1 m above ground. After 24<br />

hours the length of the resin column in the tube was measured.<br />

On l August, about 5 weeks after inoculation, exudation of resin from the point of inoculation was rec<strong>or</strong>ded on a<br />

subjective scale from 0 to 5, and used as an estimate of induced resinosis. On 14 September 1993, 81 days after inoculation<br />

all trees were felled, and tree height and height of the lowest green wh<strong>or</strong>l of branches were measured. The inoculated stem<br />

section was cut out, and brought to the lab<strong>or</strong>at<strong>or</strong>y.<br />

193


in the lab<strong>or</strong>at<strong>or</strong>y two ca. 5 mm thick cross-sectional discs were cut 20 cm inside each end of the section. The<br />

he_._wo_>d-+sapwood b<strong>or</strong>der and blue-stained areas were marked out on these two discs. Two separate measuren_ents f<strong>or</strong><br />

fung_t sc_ccess were used: (I) percent of the sapwood which had become blue-stained, and (2) percent of the disc circurn ferer_ce<br />

where the cambial area had been killed. F<strong>or</strong> both measurements, the disc having the most ft._ngal proliferation was used<br />

t(_ represen_ the tree.<br />

Above the lower of these thin discs a thicker one was cut. This ca. 3 cm disc was used f<strong>or</strong> re-isolation of' fungi from<br />

Ihe wood. Three small wood chips were taken along one radius of the disc where staining occurred, one at the front of the<br />

advancing blue=s'tain, one near the carnbium, and one in between.<br />

The data were analyzed using the Minitab statistical system (Ryan et al. 1992).<br />

RESUUFS<br />

Dt.Jrin_4the period late October 1992 to late June 1993 precipitation at As amounted to ca. 400 ram, and a further 200<br />

mm f:cll fron_ late June (inoculation) till inid-September (felling). Although small quantities of water must have leaked<br />

{hrot_gh the ground cover and later the roof, the experimental trees were still deprived of most of <strong>this</strong> precipitation. The<br />

spring of 1993 was drier than n<strong>or</strong>mal; only 124 mm felt during the period 1 March - 30 June, as opposed to a n<strong>or</strong>mal of 2 t 5<br />

mm_ N<strong>or</strong>mal annual precipitation is ca. 800 ram.<br />

Measurement of pre-dawn xylem water potentials in late June demonstrated significant differences between stressed<br />

and control trees in both clones (Fig. 1). In Clone 194 the WP values of stressed trees ranged between 0.75 and 0.95 MPa as<br />

compared to 0.35..o0.50 MPa in the control trees. No stressed tree had as high a WP as any of the controls. Clone 582<br />

exhibited les_ difference in WP between the two categ<strong>or</strong>ies of trees, and one sheltered tree (#6) was no m<strong>or</strong>e stressed than<br />

three of the control trees. Trees of the same clones at the top and bott<strong>or</strong>n of the slope showed no clear differences in WP<br />

(rest_It nol shown).<br />

On 29 June measurements indicated that stressed trees of Clone 194 yielded less constitutive resin in 24 hours than<br />

controls (p=0.0()2), but on 6 July no difference could be detected. In Clone 582 such differences vere absent on both occa-<br />

sions. Subjective estimates of surfTxce exudation of resin from the points of inoculation gave a rating that was about three<br />

lime higher in stressed trees of both clones than in c<strong>or</strong>responding controls (p=0,0(X)).<br />

LIpon felling on 14 September, no tree had lost its foliage. Several trees of Clone 194 had a yellow tint to their<br />

fi_liage, but <strong>this</strong> occurred both in stressed and control trees.<br />

The cross-sectional discs showed sapwood blue-staining varying from 0 to 100 percent (Fig. 2). Staining was much<br />

m<strong>or</strong>e pronounced in controls than in drought stressed trees of Clone 194. The stressed tree that had the most fungal invasion<br />

had mucf_ tess s{ain than the least affected control (p=O.000). In Clone 582, on the other hand, there was no difference<br />

between stressed and control trees. This measure of fungal success coincided with percentage of the disc periphery where the<br />

cambial area had been killed (Fig, 2). There was no overlap between stressed trees and controls in Clone 194 (p=0.()0()). [n<br />

Chine 582 conm_t _rees proved to be somewhat m<strong>or</strong>e affected than the stressed ones (p=0.033).<br />

O. pol<strong>or</strong>uc_m was reisolated from 10 of the 12 trees of Clone 194 and 8 out of 12 in Clone 582. Because almost 12<br />

weeks elapsed from i_oculation to felling, other fungi had also invaded the wood (cf. Solheim 1992). These were mainly<br />

Nectria spp. (f:o_.md in 14 trees), but another Ophiostoma species, O. piceae (Mtinch) H. and R Syd., occurred together wiih<br />

O. I;,o!


-0.2<br />

0<br />

5 82.<br />

_j -0,4 e a a a<br />

D4<br />

E<br />

I ® e ®<br />

-0,6 ®<br />

_4 e<br />

.4 4.3<br />

C; -0.8<br />

o<br />

.#<br />

o<br />

_4 _ .... t I 0 O 0 | . l<br />

(1)<br />

194<br />

-0 .2<br />

- 0.4 _<br />

if,<br />

3:<br />

o<br />

zx<br />

a<br />

zx<br />

_I 0<br />

-o. 6<br />

04 o o<br />

- o. 8 "<br />

@<br />

e<br />

-1 0 | I I I I I I I ....I I |<br />

1 2 3 4 5 6 7 8 9 i0 ll 12<br />

Tree number<br />

Figure 1.--Pre-dawn xylem water potential of N<strong>or</strong>way spruce trees on 23 June 1993. Trees 1-6 of both clones (194 and 582)<br />

are artificially drought stressed; trees 7-12 are controls.<br />

An earlier experiment in N<strong>or</strong>way spruce addressed possible after-effects of drought on the resistance to O. polonicum<br />

infection (Christiansen 1992). In that case, three summers of severe artificial drought did not affect the resistance of the trees<br />

during a fourth season of n<strong>or</strong>mal precipitation. Thus neither that experiment n<strong>or</strong> the present one supp<strong>or</strong>ts a general hypothesis<br />

that drought stress makes coniferous trees m<strong>or</strong>e susceptible to attack by bark beetles and their associated blue-stain fungi<br />

(cf. Christiansen et al. 1987).<br />

The drought stress was relatively mild in the present study, though possibly adequate to have influenced imp<strong>or</strong>tant<br />

physiological processes in the trees (cf. Hsiao 1973), even including gas exchange (cf. Havranek and Benecke 1978). Such<br />

changes may be thought to reduce resistance, but vascular wilts are characteristically found to be m<strong>or</strong>e severe in wet than in<br />

dry soils (Schoeneweiss 1986).<br />

A<br />

195


109 ICO<br />

I 9 4 _ _7 _ _ I00- 582 _ 92<br />

25 30<br />

._°_+m+,|_+ ...... ++.+,5_+_ .,_ o, +- ,<br />

1 2 3 4 5 6 7 8 9 I0 II 12 1 2 3 4 5 6 7 8 9 I0 II 12<br />

E I00 I00 94<br />

._ _ _y 194<br />

_ 88 _ 90 89_ 582 _ 86 88<br />

40 _'+' + i .... 2; _! ;+<br />

+_3o i+ + R > + 2+ ++{ +o ++l+ .?LI ++ .... +<br />

°" o J[Jt ......... +...... o .... m+_m......<br />

1 2 3 4 5 6 7 8 9 i0 11 12 1 2 3 4 5 6 7 8 9 I0 II ]2 13<br />

Tree nunbe_< Tree ntmber<br />

Figure 2.--Fungal proliferation in N<strong>or</strong>way spruce trees inoculated with Ophiostoma polonicum. Trees 1-6 of both clones<br />

(194 and 582) are artificially drought stressed; trees 7-12 are controls. Upper graphs: Percent blue staining of the<br />

sapwood cross-sectional area. Lower graphs: Percent cambial killing.<br />

In a field experiment in young Pinus taeda, long-lasting water deprivation resulted in pre-dawn xylem water potentials<br />

of the same magnitude as in the present study. Although resin flow was significantly reduced in the stressed pines, they<br />

were also less attacked by Dendroctonusfrontalis than the irrigated controls. The conclusion was that drought of <strong>this</strong> extent<br />

does not necessarily lower resistance (Dunn and L<strong>or</strong>io 1993).<br />

Like several other inoculation experiments in N<strong>or</strong>way spruce, the present one demonstrated that the invasive success<br />

of O. polonicum is negatively c<strong>or</strong>related with the quantities of resin that accumulate in "reaction zones" around the points of<br />

infection and ooze out from these sites (Christiansen and H<strong>or</strong>ntvedt 1983, Christiansen 1985a, Christiansen and Ericsson<br />

1986, H<strong>or</strong>ntvedt 1988, H<strong>or</strong>ntvedt and Solheim 1991). As in other conifers, <strong>this</strong> "secondary" resinosis is assumed to be due<br />

mainly to localized, induced defence reactions (Berryman 1972).<br />

During the early stages of invasion in conifers, the bark beetles encounter a first line of defence, i.e., constitutive <strong>or</strong><br />

primary resin oozing from severed ducts in the phloem. In our study there was no unequivocal difference between stressed<br />

and control trees in <strong>this</strong> character, although the first of two measurements gave higher resin yields in control trees of Clone<br />

194 than in stressed ones. However, in the end, these controls turned out to be less resistant to infection than the stressed<br />

trees. In N<strong>or</strong>way spruce, the quantities of primary resin appear to be highly variable (Christiansen 1991). Despite its<br />

apparent imp<strong>or</strong>tance f<strong>or</strong> the defence of certain pines (Cates and Alexander 1982), <strong>this</strong> constitutive resin appears to be less<br />

imp<strong>or</strong>tant f<strong>or</strong> others such as Pinus sylvestris (Schroeder 1990). It has been suggested that tree species that are repeatedly<br />

exposed to attacks by multiple generations of beetles per year may rely m<strong>or</strong>e on a constitutive defences than those which<br />

generally have only one critical period of attack (Matson and Hain 1983).<br />

F<strong>or</strong> some pine bark beetles (e.g., D. frontalis and D. brevicomis), studies suggest that drought has a negative effect on<br />

resistance, and that <strong>this</strong> effect is cumulative both on a sh<strong>or</strong>ter and a longer time scale (Craighead 1925, Miller and Keen<br />

196


1960, Kalkstein 1976). Increased susceptibility may be due to a reduced yield of constitutive resins. However, <strong>this</strong> explanation<br />

may not hold when drought is only mild (cf. Dunn and L<strong>or</strong>io 1993).<br />

It seems likely that the degree of stress to which trees are exposed will influence their resistance. Assuming that a<br />

higher level of stress would have made the trees m<strong>or</strong>e susceptible, it is relevant to ask how often it occurs in the field under<br />

natural conditions. In <strong>this</strong> study the sheltered trees were deprived (depending on their root systems) of as much as half a<br />

year's precipitation (c. 400 mm) pri<strong>or</strong> to inoculation, and a further 200 rnm were lost thereafter. Such an abrupt loss of water<br />

does not seem likely to occur under field conditions in N<strong>or</strong>way. At the n<strong>or</strong>rnat time of mass flight by I. typographus (i.e., in<br />

May), soil water reserves are not likely to be seriously depleted.<br />

In the present case it can be hypothesized that the relativley mild drought may have triggered a mobilization of some<br />

defence mechanism in Clone 194; it may have induced tile accumulation of defensive compounds <strong>or</strong> have elicited the<br />

development of defence-related cell types (V.R. Franceschi, pets. comm.). A m<strong>or</strong>e severe drought might have prevented the<br />

mobilization of such defences by critically reducing the availability of some imp<strong>or</strong>tant resources, but <strong>this</strong> remains to be<br />

proved.<br />

SUMMARY<br />

Young N<strong>or</strong>way spruce trees of two different clones were artificially deprived of precipitation water over a period of 8<br />

months, starting in late October. In late June, pre-dawn water potentials of the experimental trees were significantly lower<br />

than those of unstressed control trees, but the stress was still tlairly mild. At <strong>this</strong> time both categ<strong>or</strong>ies of trees were inoculated<br />

with a standard dose of the blue-stain fungus Ophiostoma polonicum, a usual associate of the spruce bark beetle Ips<br />

typographus. The stressed trees of one clone proved to be significantly m<strong>or</strong>e resistant to infection than unstressed controls of<br />

the same clone. The other clone showed a similar but much less pronounced trend. It is suggested that the mild stress has<br />

triggered the mobilization of some defence mechanism in the f<strong>or</strong>mer clone.<br />

ACKNOWLEDGEMENTS<br />

The study was supp<strong>or</strong>ted by a grant from the Agricultural <strong>Research</strong> Council of N<strong>or</strong>way (Contract no. 205030).<br />

Thanks are due to T<strong>or</strong>e SkrOppa and colleagues f<strong>or</strong> providing experimental trees; to Line Hamar, Heidi Knutsen, Olaug<br />

Olsen, and T<strong>or</strong>finn Smther f<strong>or</strong> field assistance; to Olaug Olsen f<strong>or</strong> fungal reisolation; to Olaug Olsen and Halv<strong>or</strong> Solheim f<strong>or</strong><br />

fungal identification; and to Halv<strong>or</strong> Solheim and Ye Hui f<strong>or</strong> critical examination of the manuscript.<br />

LITERATURE CITED<br />

BERRYMAN, A.A. 1972. Resistance of conifers to invasion by bark beetle-fungus associations. BioScience 22: 598-602.<br />

CATES, R.G. and ALEXANDER, It. 1982. Host Resistance and Susceptibility, p. 212-263. In Mitton, J.B. and Sturgeon,<br />

K.B., eds. Bark Beetles in N<strong>or</strong>th American Conifers. University of Texas Press, Austin.<br />

CHRISTIANSEN, E. 1985a. Ceratocystis polonica inoculated in N<strong>or</strong>way spruce: Blue-staining in relation to inoculum<br />

density, resinosis, and tree growth. Eur. J. F<strong>or</strong>. Pathol. 15' 160-167.<br />

CHRISTIANSEN, E. 1985b. lps/Ceratocystis-infection of N<strong>or</strong>way spruce: What is a deadly dosage? Z. Angew. Entomol.<br />

99:6-11<br />

CHRISTIANSEN, E. 199 I. Ips typographus and Ophios'toma polonicum versus N<strong>or</strong>way spruce: Joint attack and host<br />

defence, p. 321-324. In Baranchikov, Y., Mattson, W.J., Hain, F., and Payne, T., eds. F<strong>or</strong>est insect guilds: patterns of<br />

interaction with host trees. Gen. Tech. Rep. NE-153. Radn<strong>or</strong>, PA: U.S. Department of Agriculture, F<strong>or</strong>est Service:<br />

321-334.<br />

197


CHRISTIANSEN, E. 1992. After-effects of drought did not predispose young Picea abies to infection by the bark beetletransmitted<br />

blue-stain fungus Ophiostoma polonicum. Scand. J. F<strong>or</strong>. Res. 7: 557-569.<br />

CHRISTIANSEN, E. and BAKKE, A. 1988. The spruce bark beetle of Eurasia, p. 479-503. In Berryman, A.A., ed. Dynamics<br />

of F<strong>or</strong>est Insect Populations. New Y<strong>or</strong>k and London: Plenum Publishing C<strong>or</strong>p<strong>or</strong>ation.<br />

CHRISTIANSEN, E. and ERICSSON, A. 1986. Starch reserves in Picea abies in relation to defence reaction against a bark<br />

beetle transmitted blue-stain fungus, Ceratocystispolonica. Can. J. F<strong>or</strong>. Res. 16: 78-83.<br />

CHRISTIANSEN, E. and HORNTVEDT, R. 1983. Combined Ips/Ceratocystis attack on N<strong>or</strong>way spruce, and defensive<br />

mechanisms of the trees. Z. Angew. Entomol. 96:110-118.<br />

CHRISTIANSEN, E. and SOLHEIM, H. 1990. The bark beetle-associated blue-stain fungus, Ophiostoma polonicum, can<br />

kill various spruces and Douglas fir. Eur. J. F<strong>or</strong>.Pathol. 20: 436-446.<br />

CHRISTIANSEN, E., WARING, R.H., and BERRYMAN, A.A. 1987. Resistance of conifers to bark beetle attack: Searching<br />

f<strong>or</strong> general relationships. F<strong>or</strong>. Ecol. Manage. 22: 89-106.<br />

CRAIGHEAD, F.C. 1928. Interrelation of tree-killing bark beetles (Dendroctonus) and blue stains. J. F<strong>or</strong>estry 26: 886-<br />

887.<br />

DUNN, J.R and LORIO, EL. 1993. Modified water regimes affect photosynthesis, xylem water potential, cambial growth,<br />

and resistance of juvenile Pinus taeda L. to Dendroctonusfrontalis (Coleoptera: Scolytidae). Environ. Entomol. 22:<br />

948-957.<br />

FURNISS, M.M., SOLHEIM, H., and CHRISTIANSEN, E. 1990. Transmission of blue-stain fungi by Ips typographus<br />

(Coleoptera: ScoOytidae) in N<strong>or</strong>way spruce. Ann. Entomol.Soc. Am. 83: 712-716.<br />

HAVRANEK, W.M. and BENECKE, U. 1978. The influence of soil moisture on water potential, transpiration and<br />

photosythesis of conifer seedlings. Plant and Soil 49:91-103.<br />

HORNTVEDT, R. 11988. Resistance of Picea abies to Ips typographus: Tree response to monthly inoculations with<br />

Ophiostoma polonicum, a beetle transmitted blue-stain fungus. Scan. J. F<strong>or</strong>. Res. 3" 107-114.<br />

HORNTVEDT, R., CHRISTIANSEN, E., SOLHEIM, H., and WANG, S. 1983. Artificial inoculation with lps typographusassociated<br />

blue-stain fungi can kill healthy N<strong>or</strong>way spruce trees. Medd. N<strong>or</strong>. inst. skogf<strong>or</strong>sk. 38(4): 1-20.<br />

HORNTVEDT, R. and SOLHEIM, H. 1991. Pathogenicity of Ophiostoma polonicum to N<strong>or</strong>way spruce: The effect of<br />

isolate age and inoculum dose. Medd. Skogf<strong>or</strong>sk.44(4): 1-11.<br />

HSIAO, T.C. 1973. Plant responses to water stress. Ann. Rev.Plant Physiol. 24: 519-570.<br />

KALKSTEIN, L.S. 1976. Effects of climatic stress upon outbreaks of the southern pine beetle. Environ. Entomol.5: 653-<br />

658.<br />

MATSON, RA. and HAIN, F.R 1985. Host conifer defence strategies, p. 33-42. In Safranyik, L., ed. The Role of the Host<br />

in the Population Dynamics of F<strong>or</strong>est Insects. Proc. IUFRO Conf., Banff, Alberta, Canada, 4-7 Sept. 1983. Pacific<br />

F<strong>or</strong>. Res. Centre, Can. F<strong>or</strong>. Serv., Vict<strong>or</strong>ia B.C.<br />

MILLER, J.M. and KEEN, F.R 1960. Biology and control of thewestern pine beetle. U.S. Department of Agriculture,<br />

381 p.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1983. The role of host plant resistance in the colonization behavi<strong>or</strong> and ecology of<br />

bark beetles (Coleoptera: Scolytidae). Ecol. Monogr. 53: 27-49.<br />

198


RYAN, B.F., JOtNER, B.L., and RYAN, T.A., Jr. t992. Minitab Handbook. Second edition. Boston: PWS-KENT Publ.<br />

Comp.<br />

SCHOENEWEISS, D.F. t986. Water stress predisposition to disease - an overview, p. 157-174. In Ayres, EG. and Boddy,<br />

L., eds. Water, fungi and plants. Cambridge University Press.<br />

SCHROEDER, M. 1990. Duct resin flow in Scots pine in relation to attack of the bark beetle, Tomicus piniperda (L.) (Col.,<br />

Scolytidae). J. Appl. Ent. 109:t05-112.<br />

SOLHEIM, H. 1986. Species of Ophiost<strong>or</strong>nataceae isolated from Picea abies infested by the bark beetle, Ips O'pographus.<br />

N<strong>or</strong>d. J. Bot. 6: 19%207.<br />

SOLHEIM, H. 1988. Pathogenicity of some Ips 07)ographt_s-associated blue-stain fungi to N<strong>or</strong>way spruce. Medd. N<strong>or</strong>.<br />

inst. skogf<strong>or</strong>sk. 40(14): 1-11.<br />

SOLHEIM, H. t992. The early stages of fungal invasion in N<strong>or</strong>way spruce infested by the bark beetle, Ips O'pographus.<br />

Can. J. Bot. 70: 1-5.<br />

THALENHORST, W. t958. Grundztige der Populationdynamik desgrossen Fichtenb<strong>or</strong>kenk_ifers Ips typographus L.<br />

SchrReihef<strong>or</strong>stl. Fak. Univ. G6ttingen 21: 1-126.<br />

199


A[ PROACHES TO STUDYING ENVIRONMENTAL EFFECTS ON<br />

RESISTANCE OF PINUS 7AEDA L.<br />

TO DENDROCTONUS FRONTALIS Z IMM.ERMANN<br />

..... ._ ,<br />

PETER L. LORIO, JR.<br />

<strong>USDA</strong> F<strong>or</strong>est Service, Southern F<strong>or</strong>est Experiment <strong>Station</strong>, 2500 Shrevep<strong>or</strong>t Highway, Pineville, Louisiana 7 t360, USA<br />

INTRODUCTION<br />

E_vh>n_nental conditions and the genetic potential of loblolly pines, Pinus taeda L., affect <strong>or</strong> determine tree resista_ce<br />

to ;:mackby the southern pine beetle, DendroctonusJ?ontalis Zimmermann, by operating on physiological processes<br />

(Fig_ _). Er_vironment and genetic potential must operate through physiological processes to determine the quantity and<br />

quality of growdu as well as to express resistance to invasion by pathogens and bark beetles (Kramer t986). Only in <strong>this</strong> way<br />

can eiflaer e,_vironment <strong>or</strong> genetics affect growth and development at the cell <strong>or</strong> the whole tree level. Itere, I would like to<br />

focus primarily cm one maj<strong>or</strong> aspect of the environment that commonly affects the growth and development of lobtolly pines<br />

gradtheir _etative resistance to attack by the southern pine beetle; that is, the water regimes under which they may grow. It is<br />

imp<strong>or</strong>tant _:_consider the eft_ectsof environment across a range of time fi'ames, from very sh<strong>or</strong>t (diurnal <strong>or</strong> even hourly) to<br />

very h:m_(litk:time). F<strong>or</strong> example, trees growing on wet sites and in humid environments may grow rapidly and reach large<br />

size over king time frames because of prolonged wet conditions; however, in sh<strong>or</strong>t time frames of days <strong>or</strong> weeks they may be<br />

su}:_icctedto severe water deficits not evident when data are summed over long periods.<br />

ROLE OF PHYSIOLOGY IN FORESTRY<br />

GENETIC POTENTIAL ENVIRONMENTAL FACTORS<br />

Tree Improvement Atmospheric, soil,<br />

programs and biotic<br />

Silvicultural treatments<br />

PHYSIOLOGICAL PROCESSES<br />

At whole plant level<br />

At cellular level<br />

QUANTITY AND QUALITY<br />

OF GROWTH<br />

(Usually far below the possible maximum)<br />

Figure I.o--Diagram illustrating the role of physiology in f<strong>or</strong>estry. Genetic potential and environment operate through physiological<br />

processes in determining the quantity and quality of growth. Expression of resistance mechanisms of conifers<br />

to invasion by patho,,ense. _ and bark beetles is likewise governed by, these relationships (From Kramer, 1986).<br />

Ma{tson, W,J., Niemel_i, E, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest tbr pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong> Serv. Gen. Tech. Rep. NC-t83, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

2{X}


It is seldom feasible to control water regime in the practice of tbrest management as one can control the density and<br />

spacing of trees, <strong>or</strong> the sites on which trees are grown. They may be planted in various ways, on dry <strong>or</strong> wet sites, thinned to<br />

reduce competition among individuals f<strong>or</strong> light, water, and nutrients, and fertilized <strong>or</strong> pruned if desired, but in the southeastern<br />

United States there is little one can do about the great variation in rainfall through growing seasons and from year to year.<br />

Varying water regime can be a maj<strong>or</strong> fact<strong>or</strong> in tree growth and development, as well as in tree resistance to southern pine<br />

beetle attack.<br />

One may approach the study of tree responses to water regimes in a variety of ways. Commonly, a manipulative<br />

approach is chosen to establish a range of conditions that may produce measurable responses considered imp<strong>or</strong>tant to the<br />

purpose. Although it is difficult, if not impossible, to modify one aspect of the environment without altering one <strong>or</strong> m<strong>or</strong>e<br />

other fact<strong>or</strong>s, manipulative studies can be very useh_l. However, it helps considerably to consider the possible effects of<br />

prevaili ng environmental conditions and the ontogenetic stage of trees on responses to treatment.<br />

Observational studies can be very helpful, if not essential, to the development of the most useful manipulative<br />

studies. By conducting such studies, one can develop some understanding that can help in the design of better studies in<br />

which treatments are imposed. The two approaches can be complementary. F<strong>or</strong> example, (Chou 1982) emphasized the<br />

imp<strong>or</strong>tance of seasonal predisposition of host trees in planning inoculation trials f<strong>or</strong> any purpose. He found that Pinus<br />

radiata susceptibility to Diplodia pinea varied with season of the year, being high in spring-summer but low in autumnwinter.<br />

Knowledge of structural and physiological changes associated with ontogeny can be invaluable to interpretation of<br />

results of fungal inoculation studies, as indicated in studies by (Paine 1984) and (Owen et al. 1987), as well as with pinewood<br />

nematode inoculations (Myers 1986). Similarly, physiological changes that occur during ontogeny of trees may alter their<br />

susceptibility to f<strong>or</strong>est pests (Kozlowski 1969).<br />

We have conducted a number of observational studies toward discovering tree and site characteristics c<strong>or</strong>related with<br />

the initiation and recurrence of southern pine beetle infestations (L<strong>or</strong>io 1966, 1968; L<strong>or</strong>io and Hodges 1968b, 1971; Hodges<br />

and L<strong>or</strong>io 1971; L<strong>or</strong>io et al. 1972; L<strong>or</strong>io and Bennett 1974). These studies led to practical applications, such as the development<br />

of stand risk rating f<strong>or</strong> the southern pine beetle (L<strong>or</strong>io 1978, 1980; L<strong>or</strong>io and Sommers 1981a, 1981b; L<strong>or</strong>io et al. 1982;<br />

Zamoch et al. 1984; Hedden and L<strong>or</strong>io 1985). A compendium on the southern pine beetle was based to a large extent on<br />

observational studies (Thatcher et al. t980).<br />

Here, I would like to review some studies that we have conducted with loblolly pine and the southern pine beetle in<br />

attempts to determine the effect of water regime on tree resistance to beetle attack. They include both manipulative and<br />

observational studies. Contrary to our early assumption that stress in the f<strong>or</strong>m of water deficit is bad f<strong>or</strong> trees and good f<strong>or</strong><br />

bark beetles, a m<strong>or</strong>e sound assumption is that "it depends." It depends on a number of fact<strong>or</strong>s, such as tree size, growth stage<br />

<strong>or</strong> age, the timing of the deficit during the growing season, how severe the deficit becomes, how long it lasts, site conditions,<br />

tree root distribution and condition, and perhaps many other fact<strong>or</strong>s. It is especially imp<strong>or</strong>tant to remember that any such<br />

fact<strong>or</strong>s affect trees by operating through physiological processes (Kramer 1986).<br />

METHODS<br />

Early on we conducted both manipulative and observational studies to assess the effects of environmental conditions,<br />

principally soil water supply, on loblolly pine physiology and susceptibility to southern pine beetle attack. One of our<br />

manipulations was to dig trenches around groups of trees <strong>or</strong> individuals, and line them with polyethylene sheeting to restrict<br />

lateral movement of water and prevent roots of study trees from tapping soil water beyond the trench (L<strong>or</strong>io and Hodges<br />

1968b, 1977; Hodges and L<strong>or</strong>io 1969). We constructed various types of shelters that limited soil water recharge from<br />

rainfall and allowed us access to the trees f<strong>or</strong> measurements. Other methods included continuously flooding tree root<br />

systems f<strong>or</strong> long periods of time by maintaining ponded water on the site, and applying collars of solid CO 2 around the lower<br />

tree bole to inhibit the movement of water through sapwood to induce accute water deficit (Moeck et al. 198t, Stephen et al.<br />

1988). In m<strong>or</strong>e recent times we have girdled trees (Dunn and L<strong>or</strong>io 1992), sheltered root systems without trenching, and<br />

irrigated in attempts to provide wen-watered conditions to compare with the effects of sheltering (Dunn and L<strong>or</strong>io 1993).<br />

This and other studies included artificially inducing southern pine beetle attacks (L<strong>or</strong>io and Hodges 1977; Dunn and L<strong>or</strong>io<br />

1992, 1993), but in some studies, volunteer attack was relied upon (L<strong>or</strong>io and Hodges 1968b, Hodges and L<strong>or</strong>io 1969).<br />

201


Observational studies included examining site and stand conditions associated with southern pine beetle infested<br />

stands compared with noninfested stands (L<strong>or</strong>io 1968), measuring soil water regimes, depth to free water in perf<strong>or</strong>ated<br />

tubing, and tree diameter growth (L<strong>or</strong>io and Hodges 1971); determining tree rooting characteristics and distribution, and<br />

ole<strong>or</strong>esin exudation pressures in relation to micr<strong>or</strong>elief (L<strong>or</strong>io and Hodges 1968a, L<strong>or</strong>io et al. 1972); and comparing methods<br />

of measuring field techniques f<strong>or</strong> assessing the water status of trees (Hodges and L<strong>or</strong>io 1971). One study, conducted over<br />

several years, included calculation of soil water balance (Zahner and Stage 1966), measurement of cambial growth by<br />

periodically extracting xylem plugs from tree boles and counting the number of tracheids produced in the current annual ring,<br />

counting vertical resin ducts in the current annual ring and calculating their density, and periodically sampling resin flow<br />

from wounds through the years (L<strong>or</strong>io et al. 1990).<br />

RESULTS<br />

Manipulative studies to establish levels of water deficit that would severely weaken trees relative to nontreated<br />

controls (L<strong>or</strong>io and Hodges 1968b, 1977; Hodges and L<strong>or</strong>io 11969,1975) demonstrated that severe water deficits, aside from<br />

reducing growth, caused physical and chemical changes apparently associated with reduced tree resistance to beetle attack.<br />

However, it was not demonstrated that the southern pine beetle preferentially attacked trees subjected to severe water deficit,<br />

<strong>or</strong> that they required such weakened hosts tbr successful attack. On the contrary, it was clear that vig<strong>or</strong>ously growing hosts<br />

could be overcome readily by <strong>this</strong> bark beetle species, and that such hosts could produce abundant brood.<br />

An observational study of relationships among xylem water potential of twigs (Scholander et al. 1965), relative water<br />

content of needles (Weatherley 1950), ole<strong>or</strong>esin exudation pressure in the lower main bole (Vit6 1961), and sh<strong>or</strong>t-term flow<br />

of resin (Mason 1969, 1971) showed good relationships between resin pressure and the conventional measures of water<br />

status, but no relationship between resin pressure and flow (Hodges and L<strong>or</strong>io 1971). The question of a direct relationship, <strong>or</strong><br />

lack thereof, between ole<strong>or</strong>esin exudation pressure and ole<strong>or</strong>esin flow from bark wounds is discussed elsewhere (L<strong>or</strong>io in<br />

press).<br />

Recent attempts to evaluate the effects of environmental conditions on loblolly pine resistance to the southern pine<br />

beetle provide some insight about the imp<strong>or</strong>tance of environment to tree resistance. Results from over a 2-year period (1984-<br />

1985) of an observational study (L<strong>or</strong>io et al. 1990) indicated substantial differences between years in calculated soil water<br />

st<strong>or</strong>age, cumulative soil water deficits, cambial growth, density of resin ducts in the current year's annual ring, and the<br />

patterns of resin yields through the growing seasons. The results f<strong>or</strong> 1986 and 1987 are rep<strong>or</strong>ted here (Fig. 2). Clearly<br />

different water regimes prevailed from one year to the next, as did the apparent effects on cambial growth, resin duct densities,<br />

and resin yields. No treatments were applied and beetle attacks were not induced, and the apparent responses are<br />

c<strong>or</strong>relative. However, the pattern of resin yields in 1987 closely mimics a conceptual model of seasonal changes in pine<br />

resistance to beetle attack in years having soil water balance patterns similar to the long-term average (Fig. 3). Resin flow<br />

from small bark wounds is considered indicative of the relative resistance of southern pines to southern pine beetle attack<br />

(Hodges et al. 1979).<br />

A manipulative study carried out within the same stand in the spring of 1986 and the late summer of 1987 indicated a<br />

strong environmental effect preceding and during the study (Stephen et al. 1988). Another manuscript that focuses on<br />

imp<strong>or</strong>tant differences in environmental conditions, growth and development of the host trees, and southern pine beetle<br />

population levels is pending.<br />

Girdling of bark to the face of the xylem pri<strong>or</strong> to the development of latewood, and under well-watered conditions,<br />

produced considerably different responses to fungal inoculations <strong>or</strong> induced southern pine beetle attacks at 2 weeks compared<br />

to 8 weeks after girdling (Dunn and L<strong>or</strong>io 1992). Two weeks after girdling, carbohydrate status, cambial growth, resin flow,<br />

phloem moisture content, lesion f<strong>or</strong>mation, and beetle colonization differed little above <strong>or</strong> below girdles. After 8 weeks the<br />

carbohydrate status below girdles was greatly reduced, as was cambial growth (no latewood f<strong>or</strong>med). Although not statisti-<br />

cally significant, resin flow was reduced almost 50% below compared to above girdles. Phloem moisture was significantly<br />

higher below than above girdles, and beetles oviposited three times as many e_,,,s and constructed three times as much eee<br />

gallery below than above girdles. These results were obtained during a period of almost continually declining soil moisture<br />

and steadily accumulating calculated soil water deficits (Fig. 4).<br />

202


15 15<br />

u. 10 |O<br />

O<br />

N<br />

tal<br />

5 5<br />

g ao<br />

o 4o 40<br />

T<br />

I11<br />

X<br />

:0 20 20<br />

Z<br />

e_EoaO M<br />

z40 4O<br />

w<br />

a<br />

_20 20<br />

LATEWOOD<br />

_RLYWOOD<br />

EARLYWOOD ..,o./" , , __ , J , A<br />

0 _.__ .............. I ........ I _ ! l, m 1 ! I I _ 0 1 ' : : _' : : : _ ' , I I !<br />

.-- g H<br />

.= 10 10<br />

N<br />

_. B B<br />

E at 6<br />

N<br />

4 4<br />

2 2<br />

I I I I I I I I I I I I _ __! I II I II I 1 ......L ........... L__ _ _<br />

0 30 60 |10 t20 150 180 210 240 270 300 330 380 0 30 80 90 120 150 180 2t0 _N,0 270 300 330 380<br />

DAY OF'YEAR DAY OFYEAR<br />

Figure 2.--Graphs of water regimes, the course of xylem growth and development, vertical resin duct density in the<br />

current annual ring, and resin yield from standard bark wounds in 1986 (A, B, C, D) and 1987 (E, F, G, H).<br />

Daily soil water st<strong>or</strong>age (solid line) and cumulative daily water deficits (dotted line), (A, E). Tree bole cambial<br />

growth and development expressed as the number of tracheids f<strong>or</strong>med, time of transition from earlywood to<br />

latewood, and the total amount of earlywood and latewood f<strong>or</strong>med (B, F). Vertical resin duct densities in the<br />

current annual ring (C, G). Ole<strong>or</strong>esin yield over 24-hour periods (D, H). Vertical bars are standard err<strong>or</strong>s, n=l 1<br />

f<strong>or</strong> 1986, n--13 f<strong>or</strong> 1987. Water regimes, growth and development, and resin yields differ dramatically between<br />

years. Resin yields f<strong>or</strong> 1987 closely approximate the seasonal changes in resistance suggested in the conceptual<br />

model shown in Figure 3 f<strong>or</strong> resistance to southern pine beetle attack in years that have soil water regimes 203<br />

similar to that of the long-term average. Severely dry conditions in the summer of 1986 apparently not only<br />

limited carbon partitioning not only to growth, but also to secondary metabolism; e.g., ole<strong>or</strong>esin synthesis.<br />

4;


= *- _ *'- '_MODERATEWATEROEFICtTS -*<br />

Z<br />

"<<br />

_o<br />

_,a.#<br />

=<br />

EABLYWOOB LATEWO00<br />

FORMATION _- TRANSITION _ FORMATION<br />

-- VERTICALDUCTFORMATION- -- -"<br />

J F M A M J J A S 0 N D<br />

MONTHOF THEYEAR<br />

Figure 3.-,.-_,,,_A conceptual model of seasonal changes in pine resistance to southern pine beetle attack f<strong>or</strong> years that have soil<br />

water balance patterns similar to that of the long-term average. Resistance to the earliest attacking beetles is considered<br />

to be highly dependent on the potential flow of ole<strong>or</strong>esin at the wound site. (From L<strong>or</strong>io eta/. l,., 990).<br />

V<br />

=° /-'%<br />

_ \ ! \ INOCULATIONS-, ..<br />

_1 \ , SPB ATTACKj * -'"<br />

16 TREES CUT - x .<br />

X """"<br />

o 12 _,<br />

_ '_ .-" CDEF<br />

b--<br />

8<br />

Z ." X<br />

I,a,!<br />

I<br />

150 180 21 0 240 270<br />

DAY OF YEAR<br />

Figure 4.--Daily precipitation (vertical bars), calculated daily soil water st<strong>or</strong>age and cumulative daily water deficits (CDEF)<br />

tbr June-September, 1989, near Alexandria, LA. The two-level soil water st<strong>or</strong>age program of Zahner and Stage<br />

21)4 (11966)was used, _suming as sandy clay loam soil holding :20.32 cm of available water in the tree rooting zone.<br />

Dates of girdling, inoculations, and induced southern pine beetle attack are indicated. (M(_ified from Dunn and L<strong>or</strong>io<br />

1992).


Only recently have we conducted studies wilh juvenile pines {trees no_ yet into reproductive growth). One ,_ct_ study<br />

involved sheltering root systems to liinit water supply ai_d ir_igating to supplement rai__fal{ (l)uni_ am] I.,


O<br />

C<br />

Lower activity of apical meristems of the top<br />

relative to photosynthetic capacity<br />

I 1 °<br />

Excess photosynthate in the top E I1<br />

Greater translocation to other active meristems" _ _o<br />

root apical and lateral cambia _E<br />

(D<br />

and c<br />

,,,---<br />

.13 0<br />

-._ Greater differentiation of cells and tissues<br />

£:}_<br />

1. protective secondary metabolites =<br />

2. higher soluble and insoluble st<strong>or</strong>age compounds _,<br />

3. flower induction<br />

Figure 5.--Carbon partitioning among growth and developmental processes in response to changing nutrient and water<br />

supplies influence the competition and fluctuating balances between growth and differentiation processes, e.g., the<br />

various cambial meristems versus protective secondary metabolites; as well as among aboveground and belowground<br />

systems within a tree. (After G<strong>or</strong>don and Smith 1987).<br />

Voluminous literature in the area of physiological ecology and plant/herbiv<strong>or</strong>e interactions in general appears to be<br />

related closely to the growth-differentiation balance concept (Mooney and Chu 1974, Mooney et al. 1983, Bryant et al. 1983,<br />

Coley et al. 1985, Chapin et al. 1987, and others). Although the paths of research in those areas did not include consideration<br />

of Loomis's (1932) concept, I find that it provides considerable supp<strong>or</strong>t f<strong>or</strong> developments in those fields. F<strong>or</strong> example, it<br />

seems to me to be essentially in agreement with the carbon-nutrient balance hypothesis (Bryant et al. 1983, Tuomi et al.<br />

1988), but provides a broader framew<strong>or</strong>k, primarily by providing explicit consideration f<strong>or</strong> water. Several papers include<br />

consideration of the plant growth-differentiation balance concept in f<strong>or</strong>ming the<strong>or</strong>ies of evolutionary development of plant<br />

defenses against insects (Tuomi et al. 1990, Herms and Mattson 1992, Tuomi 1992).<br />

Here, I have indicated some of the research results from both early and recent studies. The m<strong>or</strong>e recent studies were<br />

carried out with consideration of plant growth-differentiation balance relationships in mind, and with considerable stimula-<br />

tion from advances in physiological ecology and studies of plant/herbiv<strong>or</strong>e interactions in general. We are continuing<br />

research in the same vein. The study with juvenile pines (Dunn and L<strong>or</strong>io 1993) has been extended to an older stand, with<br />

some improvements in techniques, and with results indicating strongly that water deficits have nonlinear effects on tree<br />

206<br />

o<br />

cO<br />

"O c<br />

,,t


esistance to beetle attack. Another study is in progress, once again with juvenile pines, in a plantation subjected to thinning<br />

and fertilization. We will be continuing w<strong>or</strong>k in that direction. At <strong>this</strong> time preliminary results indicate that the sh<strong>or</strong>t-term<br />

effects of thinning and fertilization are to enhance carbon partitioning to growth and to reduce carbon committed to secondary<br />

metabolism; e.g., resin synthesis.<br />

SUMMARY<br />

There are a number of ways to approach the problem of assessing the effects of environmental conditions, such as<br />

water regime, on tree physiological responses and resistance to bark beetle attack. It helps to keep in mind that environmental<br />

fact<strong>or</strong>s operate through physiological processes (Fig. 1, and Kramer 1986), and that there are concepts, such as plant<br />

growth-differentiation balance (Loomis 1932) and carbon-nutrient balance (Bryant et al. 1983; Tuomi 1992; Tuomi et al.<br />

1988, 1990), that can provide bases f<strong>or</strong> f<strong>or</strong>ming testable hypotheses. It is especially imp<strong>or</strong>tant to know as much as possible<br />

about the host tree and the specific bark beetle of interest. Observational studies are particularly imp<strong>or</strong>tant because they can<br />

provide a baseline f<strong>or</strong> interpreting the results of manipulative studies. Whenever feasible, it is especially helpful in manipulative<br />

studies to characterize environmental conditions bef<strong>or</strong>e and during a study. Because physiological changes that occur<br />

during ontogeny of trees can alter their susceptibility to herbiv<strong>or</strong>es, knowledge of tree stage of growth and development can<br />

be especially helpful.<br />

ACKNOWLEDGMENTS<br />

Cooperative Agreements (19-86-036 and 19-87-20) between the Mississippi Agricultural and F<strong>or</strong>estry Experiment<br />

<strong>Station</strong> and the <strong>USDA</strong> F<strong>or</strong>est Service, Southern F<strong>or</strong>est Experiment <strong>Station</strong>, contributed in part to results rep<strong>or</strong>ted here.<br />

' LITERATURE CITED<br />

BRYANT, J.P., CHAPIN, F.S., III, and KLEIN, D.R. 1983. Carbon/nutrient balance of b<strong>or</strong>eal plants in relation to vertebrate<br />

herbiv<strong>or</strong>y. Oikos 40: 357-368.<br />

CHAPIN, ES., III, BLOOM, A.J., FIELD, C.B., and WARING, R.H. 1987. Plant responses to multiple environmental<br />

fact<strong>or</strong>s. BioScience 37: 49-57.<br />

CHOU, C.K.S. 1982. Susceptibility of Pinus radiata seedlings to infection by Diplodia pinea as affected by pre-inoculation<br />

conditions. New Zealand J. F<strong>or</strong>. Sci. 438-441:<br />

COLEY, P.D., BRYANT, J.P., and CHAPIN, F.S., III. 1985. Resource availability and plant antiherbiv<strong>or</strong>e defense. Science<br />

230: 895-899.<br />

DUNN, J.P. and LORIO, P.L., Jr. 1992. Effects of bark girdling on carbohydrate supply and resistance of loblolly pine to<br />

southern pine beetle (Dendroctonusfrontalis Zimm.) attack. F<strong>or</strong>. Ecol. and Manage. 50:317-330.<br />

DUNN, J.P. and LORIO, P.L., Jr. 1993. Modified water regimes affect photosynthesis, xylem water potential, cambial<br />

growth, and resistance of juvenile Pinus taeda L. to Dendroctonusfrontalis (Coleoptera: Scolytidae). Environ.<br />

Entomol. 22(5): 948-957.<br />

GORDON, J.C. and SMITH, W.H. 1987. Tree roots and microbes: a high pri<strong>or</strong>ity f<strong>or</strong> high technology, p. 423-432. In<br />

Boersma, L.L., et al., eds. Future Developments in Soil Science <strong>Research</strong>. Soil Science Society of America, Madison,<br />

WI.<br />

HEDDEN, R.L. and LORIO, RL., Jr. 1985. Rating stand susceptibility to southern pine beetle attack on national f<strong>or</strong>ests in<br />

the Gulf Coastal Plain. Res. Pap. SO-221. Asheville, NC: U.S. Department of Agriculture, F<strong>or</strong>est Service, Southern<br />

<strong>Research</strong> <strong>Station</strong>. 5 p.<br />

207


HERMS, D.A. AND MATTSON, W.J. 1992. The dilemma of plants: to grow <strong>or</strong> defend. Quart. Rev. Biol. 67: 283-335.<br />

HODGES, J.D. and LORIO, EL., Jr. 1969. Carbohydrate and nitrogen fractions of the inner bark of loblolly pines under<br />

moisture stress. Can. J. Bot. 47: 1651-1657.<br />

HODGES, J.D. and LORIO, EL., Jr. I971. Comparison of field techniques f<strong>or</strong> measuring moisture stress in large loblolly<br />

pines. F<strong>or</strong>. Sci. 17: 220-223.<br />

HODGES, J.D. and LORIO, EL., Jr. 1975. Moisture stress and composition of xylem ole<strong>or</strong>esin in loblolly pine. F<strong>or</strong>. Sci. 22:<br />

283-290.<br />

HODGES, J.D., ELAM, W.W., WATSON, W.E, and NEBEKER, T.E. 1979. Ole<strong>or</strong>esin characteristics and susceptibility of<br />

four southern pines to southern pine beetle (Coleoptera: Scolytidae) attacks. Can. Entomol. 111:8 89-896.<br />

KLEBS, G. 1903. WiIlkurliche Entwickelungsanderungen bei Pflanzen. G. Fisher, Jena.<br />

KLEBS, G. 1910. Alterations in the development and f<strong>or</strong>ms of plants as a result of environment. Proceedings of the Royal<br />

Society of London 82B: 547-548.<br />

KOZLOWSKI, T.T. 1969. Tree physiology and f<strong>or</strong>est pests. J. F<strong>or</strong>. 67:118-123.<br />

KRAMER, P.J. 1986. The role of physiology in f<strong>or</strong>estry. Tree Physiol. 2: 1-16.<br />

LOOMIS, W.E. 1932. Growth-differentiation balance vs carbohydrate-nitrogen ratio. Proceedings of the American Society<br />

f<strong>or</strong> H<strong>or</strong>ticultural Science 29: 240-245.<br />

LOOMIS, W.E. 1953. Growth c<strong>or</strong>relation, p. 197-217. In Loomis, W.E., ed. Growth and Differentiation in Plants. Iowa<br />

State College Press, Ames, IA.<br />

LOOMIS, R.S., LUO, Y., and KOOMAN, E 1990. Integration of activity in the higher plant, p. 105-124. In Rabbinge, R. et<br />

al., eds. The<strong>or</strong>etical Production Ecology: Reflections and Perspectives. Pudoc, Wageningen, The Netherlands.<br />

LORIO, EL., Jr. 1966. Phytophth<strong>or</strong>a cinnamomi and Pythium species associated with loblolly pine decline in Louisiana.<br />

Plant Dis. Rep. 50: 596-597.<br />

LORIO, EL., Jr. 1968. Soil and stand conditions related to southern pine beetle activity in Hardin County, Texas. J. Econ.<br />

Entomol. 61: 565-566.<br />

LORIO, EL., Jr. 1978. Developing stand risk classes f<strong>or</strong> the southern pine beetle. Res. Pap. SO-144. Asheville, NC: U.S.<br />

Department of Agriculture, F<strong>or</strong>est Service, Southern <strong>Research</strong> <strong>Station</strong>. 9 p.<br />

LORIO, EL., Jr. 1980. Loblolly pine stocking levels affect potential f<strong>or</strong> southern pine beetle infestation. South. J. Appl. F<strong>or</strong>.<br />

4: 259-265.<br />

LORIO, EL., Jr. 1986. Growth-differentiation balance: a basis f<strong>or</strong> understanding southern pine beetle-tree interactions. F<strong>or</strong>.<br />

Ecol. and Manage. 14: 259-273.<br />

LORIO, EL., Jr. 1993. The relationship of ole<strong>or</strong>esin exudation pressure (<strong>or</strong> lack thereof) to flow from wounds. J. Sustain.<br />

F<strong>or</strong>. (In Press).<br />

LORIO, EL., Jr. and BENNETT, W.H. 1974. Recurring southern pine beetle infestations near Oakdale, Louisiana. Res.<br />

Pap. SO-95. Asheville, NC: U.S. Department of Agriculture, F<strong>or</strong>est Service, Southern <strong>Research</strong> <strong>Station</strong>. 6 p.<br />

LORIO, EL., Jr. and HODGES, J.D. I968a. Microsite effects on ole<strong>or</strong>esin exudation pressure of large loblolly pines.<br />

Ecology 49(6): 1207-1210.<br />

2O8


LORIO, EL., Jr'. and HODGES, J.D. 1968b. Ole<strong>or</strong>esin exudation pressure and relative water content of inner bark as<br />

indicat<strong>or</strong>s of moisture stress in loblolly pines. F<strong>or</strong>. Sci. 14(4): 393-398.<br />

LORIO, RL., Jr. and HODGES, J.D. 1971. Micr<strong>or</strong>elief, soil water regime, and loblolly pine growth on a wet, mounded site.<br />

Soil Science Society of America Proceedings 35: 795--800.<br />

LORIO, RL., Jr. and HODGES, J.D. 1977. Tree water status affects induced southern pine beetle attack and brood production.<br />

Res. Pap. SO-135. New Orleans, LA: U.S. Department of Agriculture, F<strong>or</strong>est Service, Southern F<strong>or</strong>est Experiment<br />

<strong>Station</strong>. 7 p.<br />

LORIO, RL., Jr. and HODGES, J.D. 1985. The<strong>or</strong>ies of interactions among bark beetles, associated micro<strong>or</strong>ganisms, and<br />

host trees, p. 485-492. In Shoulders, E., ed. Proceedings of the 3rd Biennial Southern Silvicultural <strong>Research</strong> Conference.<br />

Gen. Tech. Rep. SO-54. New Orleans, LA: U.S. Department of Agriculture, F<strong>or</strong>est Service, Southern F<strong>or</strong>est<br />

Experiment <strong>Station</strong>.<br />

LORIO, RL., Jr. and SOMMERS, R.A. 1981a. Central Louisiana, p. 23-39. In Coster, J.E. and Searcy, J.L., eds. Site, Stand,<br />

and Host Characteristics of Southern Pine Beetle Infestations. Tech. Bull 1612. Pineville, LA: U.S. Department of<br />

Agriculture, F<strong>or</strong>est Service. Combined F<strong>or</strong>est Pest <strong>Research</strong> and Development Program.<br />

LORIO, RL., Jr. and SOMMERS, R.A. 1981b. Use of available resource data to rate stands :f<strong>or</strong> southern pine beetle risk, p.<br />

75-78. In Hedden, R.L., Barras, S.J., and Coster, J.E., tech. co<strong>or</strong>ds. Proceedings of Symposium Hazard-Rating<br />

Systems in F<strong>or</strong>est Pest Management. Gen. Tech. Rep. WO-27. New Oreleans, LA: U.S. Department of Agriculture,<br />

F<strong>or</strong>est Service, Southern F<strong>or</strong>est Experiment <strong>Station</strong>.<br />

LORIO, RL., Jr., HOWE, V.K., and MARTIN, C.N. 1972. Loblolly pine rooting varies with micr<strong>or</strong>elief on wet sites.<br />

Ecology 53:1134-1140.<br />

LORIO, P L., Jr., MASON, G.N., and AUTRY, G.L. 1982. Stand risk rating f<strong>or</strong> the southern pine beetle: integrating pest<br />

management with f<strong>or</strong>est management. J. F<strong>or</strong>. 80:212-214.<br />

LORIO, RL., Jr., SOMMERS, R.A., BLANCHE, C.A., HODGES, J.D., and NEBEKER, T.E. 1990. Modeling pine resistance<br />

to bark beetles based on growth and differentiation balance principles, p. 402-409. In Dixon, R.K., Meldahl,<br />

R.S., Ruark, G.A., and Warren, W.G., eds. Process Modeling of F<strong>or</strong>est Growth Responses to Environmental Stress.<br />

Timber Press, P<strong>or</strong>tland, OR.<br />

MASON, R.R. 1969. A simple technique f<strong>or</strong> measuring ole<strong>or</strong>esin exudation flow in pine. F<strong>or</strong>. Sci. 15: 56-57.<br />

MASON, R.R. 1971. Soil moisture and stand density affect ole<strong>or</strong>esin exudation flow in a loblolly pine plantation. F<strong>or</strong>. Sci.<br />

17: 170-177.<br />

MOECK, H.A., WOOD, D.L,. and LINDAHL, K.Q., Jr. 1981. Host selection behavi<strong>or</strong> of bark beetles (Coleoptera:<br />

Scolytidae) attacking Pinus ponderosa, with special emphasis on the western pine beetle, Dendroctonus brevicomis.<br />

J. Chem. Ecol. 7(1): 49-83.<br />

MOONEY, H.A. and CHU, C. 1974. Seasonal carbon allocation in Heteromeles arbutifolia, a Calif<strong>or</strong>nia evergreen shrub.<br />

Oecologia 14: 295-306.<br />

MOONEY, H.A., GULMON, S.L., and JOHNSON, N.D. 1983. Physiological constraints on plant chemical defenses, p. 21-<br />

36. In Hedlin, RA., ed. Plant Resistance to Insects. American Chemical Society, Washington, D.C.<br />

MYERS, R.F. 1986. Cambium destruction in conifers causedby pinewood nematodes. J. Nematol. 18: 398-402.<br />

OWEN, D.R., LINDAHL, K.Q., WOOD, D.L., and PARMETER, J.R. 1987. Pathogenicity of fungi isolated from<br />

Dendroctonus valens, D. brevicomis, and D. ponderosae to ponderosa pine seedlings. Phytopathology 77: 631-636.<br />

209


PAINE, "I.D. 1984. Seasonal response of ponderosa pine to inoculation of the mycangial fungi isolated from the western<br />

pine beetle. Can. J. Bot. 62:551-555.<br />

SCHOLANDER, RE, HAMMEL, H.T., BRADSTREET, E.D., and HEMMINGSEN, E.A. 1965. Sap pressure in vascular<br />

plants. Science 148: 339-346.<br />

STEPHEN, F.M., LIH M.R, PAINE, T.D., and WALLIS G.W. 1988. Using acute stress to modify tree resistance: impact on<br />

within-tree southern pine beetle populations, p. 105-119. In Payne, T.L. and Saarenmaa, H., eds. Integrated Control<br />

of Scolytid Bark Beetles. Proceedings of the IUFRO W<strong>or</strong>king Party and XVII International Congress of Entomology<br />

Symposium, Vancouver, B.C., Canada, July 4, 1988. Virginia Polytechnic Institute and State University.<br />

THATCHER, R.C., SEARCY, J.L., COSTER, J.E., and HERTER, G.D. (eds.). 1980. The southern pine beetle. Tech. Bull.<br />

1631. U.S. Department of Agriculture, F<strong>or</strong>est Service, and Science and Education Administration. 267 p.<br />

TREWAVAS, A. 1985. A pivotal role f<strong>or</strong> nitrate and leaf growth in plant development, p. 77-91. In Baker, N.K., et al., eds.<br />

Control of Leaf Growth. Cambridge University Press, Cambridge.<br />

TUOMI, J. 1992. Toward integration of plant defence the<strong>or</strong>ies. Trends in Ecol. and Evol. 7: 365-367.<br />

TUOMI, J., NIEMELA, R, and SIREN, S. 1990. The Panglossian paradigm and delayed inducible accumulation of foliar<br />

phenolics in mountain birch. Oikos 59: 399-410.<br />

TUOMI, J., NIEMELA, R, CHAPIN, F.S., III, BRYANT, J.R, and SIREN, S. 1988. Defensive responses of trees in relation<br />

to their carbon/nutrient balance, p. 57-72. In Mattson, W.J., Levieux, J., and Bernard-Dagan, C., eds. Mechanisms of<br />

Woody Plant Defenses Against Insects. Springer-Verlag, New Y<strong>or</strong>k.<br />

VITE, J.R 1961. The influence of water supply on ole<strong>or</strong>esin exudation pressure and resistance to bark beetle attack in Pinus<br />

ponderosa. Boyce Thompson Inst Contrib 21 : 37-66.<br />

WEATHERLEY, RE. 1950. Studies in the water relations of the cotton plant. I. The field measurement of water deficits in<br />

leaves. New Phytologist 49: 81-97.<br />

ZAHNER, R. and STAGE, A.R. 1966. A procedure f<strong>or</strong> calculating daily moisture stress and its utility in regressions of tree<br />

growth on weather. Ecology 47: 64-74.<br />

ZARNOCH, S.J., LORIO, RL., Jr., and SOMMERS, R.A. 1984. A logistic model f<strong>or</strong> southern pine beetle stand risk rating<br />

in central Louisiana. J. Ge<strong>or</strong>gia Entomol. Soc. 19: 168-175.<br />

210


EFFECTS OF ROOT INHABITING INSECT-FUNGAL COMPLEXES<br />

ON ASPECTS OF TREE RESISTANCE TO BARK BEETLES<br />

KENNETH E RAFFA and KIER D. KLEPZIG _<br />

Department of Entomology, University of Wisconsin, Madison, Wisconsin 53706, USA<br />

INTRODUCTION<br />

F<strong>or</strong>est declines have proven difficult to address, because they involve complex interactions in which the key processes<br />

are only po<strong>or</strong>ly understood (Houston 1981, Witcosky and Hansen 1985, Witcosky et al. 1986, Klein and Perkins 1988,<br />

Mueller-Dombois 1988, Huettl et al. 1990, Manion 1991, Saxe 1993, Woodcock et al. 1993). Without a basic understanding<br />

of the underlying mechanisms, it is difficult to relate field observations to various management practices, anthropogenic <strong>or</strong><br />

environmental stresses, <strong>or</strong> natural successional events. One of our maj<strong>or</strong> inf<strong>or</strong>mation gaps concerns the interactions among<br />

different insect and fungal feeding guilds that both respond and contribute to plant stress physiology. We have been addressing<br />

<strong>this</strong> question in declining red pine, Pinus resinosa, stands in the Great Lakes region of N<strong>or</strong>th America. Our objective is to<br />

link field c<strong>or</strong>relations with underlying physiological mechanisms.<br />

Insects and Pathogens Associated With Declining Red Pine Stands<br />

Red Pine Decline Disease is characterized by a progressive deteri<strong>or</strong>ation and death of mature plantation trees.<br />

Affected p<strong>or</strong>tions of stands are highly localized, and m<strong>or</strong>tality progresses radially from an epicenter (Klepzig et al. 1991).<br />

No living trees remain in <strong>this</strong> epicenter, and the resulting gap is colonized by herbaceous and early succession vegetation.<br />

Trees killed during the current growing season ring <strong>this</strong> zone. Neighb<strong>or</strong>ing trees show reduced radial and crown growth,<br />

whereas trees farther into the stand show no symptoms. Each year, these zones expand. The killed trees become windblown,<br />

the low vig<strong>or</strong> trees are killed, and additional trees show signs of stress. This syndrome can be characterized at the betweenstand,<br />

within-stand, and within-tree levels. In a 3-year evaluation of 19 sites at 13 locations in Wisconsin, Klepzig et al.<br />

(1991) found that the most consistent between-stand c<strong>or</strong>relate with Red Pine Decline was the abundance of root- and lowerstem<br />

infesting beetles (Table 1). These included three curculionid and two scolytid species. Several other scolytids and<br />

curculionids were also present in high numbers, but their densities did not consistently differ between declining and asymptomatic<br />

plantations of the same age categ<strong>or</strong>y. Likewise, there were no strong differences associated with abiotic soil fact<strong>or</strong>s.<br />

The beetles associated with Red Pine Decline differ in their within-stand spatial distributions relative to the margin of<br />

tree m<strong>or</strong>tality. F<strong>or</strong> example, four beetles occur at higher densities in declining than asymptomatic stands. However, two of<br />

these show no significant within-stand distribution patterns, one occurs m<strong>or</strong>e commonly near the zone of stressed trees, and<br />

one occurs at higher densities in regions of the stand not yet showing visible symptoms (Table 1). Interestingly, the red<br />

turpentine beetle, D. valens Le Conte, occurs at higher densities in the asymptomatic than symptomatic p<strong>or</strong>tions of declining<br />

stands, yet does not show statistically significant between-stand differences.<br />

The subc<strong>or</strong>tical belowground beetle guild inhabiting Wisconsin pines shows a high degree of niche partitioning. This<br />

separation among closely related beetles is based on host tissue tree vig<strong>or</strong>, olfact<strong>or</strong>y behavi<strong>or</strong>, and symbiotic relationships<br />

(Raffa and Hunt 1989, Rieske and Raffa 1990, 1991, Raffa and Klepzig 1992, Hunt et al. 1993). F<strong>or</strong> example, H. p<strong>or</strong>culus<br />

1Currentaddress: Department of Plant and Soil Science, Southern University and A & M College, Baton Rouge, Louisiana 70813<br />

USA.<br />

Mattson, W.J., Niemel_i, R, and Rousi, M., eds. 11996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

211


Table 1.--Root- and lower-stem infesting beetles associated with Red Pine Decline Disease (From Klepzig et al. 1991).<br />

A. Species occurring at higher densities in declining than asymptomatic stands<br />

Hylobius radicis (Curculionidae)<br />

Hylobius pales (Curculionidae)<br />

Pachylobius piciv<strong>or</strong>us (Curculionidae)<br />

Hylastes p<strong>or</strong>culus (Scolytidae)<br />

B. Species occurring at higher densities in asymptomatic p<strong>or</strong>tions of declining stands<br />

Dendpvctonus vatens (Curculionidae)<br />

tlylastes p<strong>or</strong>culus (Scolytidae)<br />

C. Species occurring at higher densities in symptomatic p<strong>or</strong>tions of declining stands<br />

Pachylobius piciv<strong>or</strong>us (Curculion idae)<br />

and H. rhyzophagus colonize lateral roots; H. pales, H. radicis, and P.piciv<strong>or</strong>us colonize the root collar region; and D. vaIens<br />

enter trees just above the soil line. M<strong>or</strong>eover, beetle behavi<strong>or</strong>al responses to volatiles associated with stress (Kimmerer and<br />

Kozlowski 1982) reflect differences in preferred host physiology. F<strong>or</strong> example, H. pales and P. piciv<strong>or</strong>us differ in their<br />

relative responses to various ratios of terpene-ethanol mixtures (Rieske and Raffa 1991). Each of the beetles in Table 1 is<br />

closely associated with Ophiotomatales fungi. The scolytids are primarily associated with Leptographium terebrantis Barras<br />

and Perry, and the curculionids are primarily associated with Leptographium procerum (Kendrick) Wingfield (Klepzig et al.<br />

1991). These fungi are moderately pathogenic to P. resinosa but cannot kill healthy trees (Raffa and Smalley 1988a, Klepzig<br />

et al. 1994a). The rates of association between these beetles and fungi average around 70%, and the vect<strong>or</strong> efficiencies are<br />

around 50% (Klepzig et al. 1994b).<br />

Colonization by the beetles shown in Table 1 and their associated fungi does not kill host trees. A survey of living red<br />

pines revealed that a high percentage of the trees peripheral to the margin of tree m<strong>or</strong>tality were infested with D. valens, H.<br />

radicis, <strong>or</strong> both (H. pomulus was not sampled). Such trees can remain alive f<strong>or</strong> several years. In most cases, however, these<br />

trees are subsequently killed by the stem colonizing bark beetle lps pini (Say), and its associated fungus Ophiostoma ips<br />

(Rumb.) Nannf (Fig. 1). All of the dead, but none of the living, trees were infested with I. pini. Thus, the I. pini - O. ips<br />

complex appears to be the ultimate cause of tree death, but root infect:ion appears to be a predisposing fact<strong>or</strong>. I. pini appears.<br />

to be highly opp<strong>or</strong>tunistic. Despite its strong association with individual tree death, between-stand population densities were<br />

not associated with stand condition.<br />

The above view is supp<strong>or</strong>ted by our observations of a declining stand over a 5-year period (Fig. 2). The percentage<br />

of trees killed by I. pini that showed pri<strong>or</strong> colonization by H. radicis averaged 86.8 + 9.7(SD%). The percentage of trees<br />

killed by I. pini that showed pri<strong>or</strong> colonization by D. valens averaged 19.4 + 12.4 (SD%). Most of the trees infested by one<br />

of these root colonizing species were also infested by the other. Interestingly, the lowest prop<strong>or</strong>tion of trees killed by L pini<br />

that had previously been infested by root insects occurred during 1988, a year of severe drought (H. radicis: 72.7%; D.<br />

valens: 8.2%). Belowground excavations of five declining and control stands revealed a similar, but even m<strong>or</strong>e striking,<br />

trend (Klepzig et al. 1991). In declining stands, fungal staining and root deteri<strong>or</strong>ation proceeded in advance of the insect<br />

vect<strong>or</strong>s and any aboveground symptoms. Significant root infection occurs 10-15m beyond the margin of killed trees. In<br />

control stands, and in the p<strong>or</strong>tions of symptomatic stands far beyond the stressed trees, root infection levels are negligible.<br />

Leptographium fungi progress through root grafts, which are extensive throughout mature P resinosa plantations. The extent<br />

of grafting is so common in these plantations that the phloem tissue of stumps can remain viable f<strong>or</strong> several years (Klepzig et<br />

af. 1994a). The exact mechanism by which Leptographium fungi progress through root grafts into healthy tissue remains<br />

unknown, however.<br />

212


OCCURRENCE OF ROOT AND STEM SUBCORTICAL<br />

BEETLES JN LmVINGRED PINES<br />

NONE i<br />

D. VALENS<br />

HYLOBIUS<br />

I, PIN_ N =<br />

184<br />

_----T------__----T------<br />

0 10 20 30 40 50 60<br />

% iNFESTED<br />

OCCURRENCE OF ROOT AND STEM SUBCORTICAL<br />

BEETLES IN KILLED RED PINES<br />

NONE<br />

,,_<br />

D. VALENS _<br />

HYLOBIUS<br />

/ N = 229<br />

I. PINi i<br />

..... "" l • I " I "'" .... _ ............ " .......... i<br />

0 20 40 60 80 100<br />

% INFESTED<br />

Figure 1.--Association of root- and stem-feeding subc<strong>or</strong>tical beetles with living (above) and dead (below) red pines in<br />

declining plantations.<br />

213


.,_.o,4e,., HQ,,N,.,_<br />

tn 8 O- H. radicis<br />

UJ I00- __ K,,.,._<br />

IJl.! ",Xx,, "<br />

_" 60- xx _ 1987<br />

t_<br />

I.U ........X ........ 1992<br />

t.t) 40-<br />

o_ 20-<br />

__<br />

(D O tt') It)<br />

V v.- £Xl (Xl<br />

, i A<br />

t,.O 'r'y-,,.<br />

100 - x ......... -x_<br />

a.... .----_',::,,.. . .<br />

"_,:., los pm/<br />

t,_ 80" --'-"-O - 1987<br />

" ", .... 0 .... 1989<br />

(/") .... 4_.... 1990<br />

m 40" ,,. \ x • "...<br />

" N \ \""-.:"-.. ---o---199_<br />

z N-, \ 'u. "'-""<br />

'- .,,.. -...::-,.,.,<br />

"---.-,,,,-..:.,:...<br />

0 _ ......... "_'_"--<br />

(D<br />

V<br />

0<br />

'rtO<br />

¢q<br />

,<br />

LO<br />

CXl<br />

A<br />

CO "--<br />

,,I-m<br />

DISTANCE FROM 1986 MARGIN OF TREE MORTALITY (m)<br />

Figure 2.--Progression of infestation in a declining red pine stand. Infestation rates of the pine root collar weevil, Hylobius<br />

radicis (above) and the pine engraver, lois pini (below) relative to the <strong>or</strong>iginally rec<strong>or</strong>ded margin of killed trees are<br />

shown.<br />

214


]Predisposition of Root-infected Trees to Stem Colonizing Insect-fungal Complexes<br />

Based on these observations, we hypothesized that pri<strong>or</strong> infection by root- and lower-stem colonizing insects and<br />

fungi reduces P. resinosa resistance against L pini - 0. ips complexes. Testing <strong>this</strong> hypothesis requires both an understanding<br />

of red pine defense mechanisms against subc<strong>or</strong>tical invasion, and comparisons of the key parameters between root-infected<br />

and non-root-infected trees.<br />

The response of P resinosa to invasion by biotic agents involves the secretion of pref<strong>or</strong>med allelochemicals from<br />

severed resin ducts, and the rapid f<strong>or</strong>mation of a necrotic lesion around, and accumulation of additional allelochemicals at,<br />

the entrance site (Raffa and Smalley 1988b, 1994). These allelochemicals are comprised primarily of monoterpenes and<br />

phenolics. This is a rather generalized response that can be elicited by multiple agents and is characterized m<strong>or</strong>e by quantitative<br />

than qualitative allelochemica[ changes. It resembles those rep<strong>or</strong>ted f<strong>or</strong> other conifer-scolytid-Ophiostoma systems (e.g.,<br />

Berryman 1972; Hodges et al. 1979; Cates and Alexander 1982; Raffa and Berryman 1982a,b; Christensen and H<strong>or</strong>ntvedt<br />

1983; L<strong>or</strong>io t986, 1993; Paine and Stephen 1987a,b; Lieutier and Berryman 1988; Nebeker et al. 1993; Raffa et al. 1993).<br />

Both constitutive and reaction allelochemicals are imp<strong>or</strong>tant to tree defense. However, allelochemical concentrations are<br />

higher in reaction tissue, and they have higher inhibit<strong>or</strong>y effects on fungal germination and growth and on insect feeding and<br />

survival (Klepzig et al. 1994d). F<strong>or</strong> example, I. pini adults strongly prefer media amended with extracts <strong>or</strong> simulated<br />

monoterpene doses from constitutive over reaction phloem tissue. Likewise, the monoterpene levels present in constitutive<br />

stem phloem tissue can kilt 60% of adult L pini within 2 days, but monoterpene concentrations present only 3 days after<br />

initiation of an induced response are sufficient to kill 90% of the beetles (Fig. 3) (Raffa and Smalley 1994). Without effective<br />

induced responses, the surviving beetles would probably attract enough additional beetles to deplete host resins and overwhelm<br />

tree resistance; without high constitutive levels, trees would be unable to delay aggregation long enough f<strong>or</strong> induced<br />

reactions to be mobilized (Raffa and Berryman 1982a, 1983a,b, 1987; Raffa et al. 1993).<br />

E _00 100<br />

c._ f,q<br />

c:_ 60 80 >-<br />

._J<br />

L) ._<br />

z 60 60 __<br />

O<br />

40 40<br />

o<br />

2:<br />

Ill TOXICITY<br />

__J MONOTERPENE CONO<br />

Z LU<br />

D_<br />

o:<br />

UJ<br />

I--<br />

0<br />

z<br />

C)<br />

20<br />

0<br />

20<br />

0<br />

D<br />

LU<br />

0:3<br />

3<br />

DAYS AFTER ATTACK<br />

Figure 3.--Effects of constitutive and induced monoterpene concentrations on survival of Ips pini adults. Monoterpene<br />

concentrations in mature red pines from days 0 (pre inoculation) to 3 following inoculation with O. ips are shown in<br />

f<strong>or</strong>eground. M<strong>or</strong>tality rate to adult I. pini f<strong>or</strong> each concentration of synthetic monoterpene in lab<strong>or</strong>at<strong>or</strong>y assay is<br />

shown in background.<br />

215


We conducted a series of comparisons between root-infected and asymptomatic trees f<strong>or</strong> a range of parameters<br />

associated with host susceptibility to I. pini - O. ips complexes. These comparisons included field assays at the whole tree <strong>or</strong><br />

controlled bait level, and lab<strong>or</strong>at<strong>or</strong>y assays involving application of controlled treatments. Both groups of assays involved<br />

behavi<strong>or</strong>al tests with I. pini, and tree physiological responses to simulated I. pini - O. ips attack. The results are summarized<br />

in Table 2.<br />

Behavi<strong>or</strong>al assays evaluated three phases of scolytid <strong>or</strong>ientation to host trees: initial landing, host entry, and gallery<br />

f<strong>or</strong>mation (Wood 1972, Elkinton and Wood 1980, Raffa and Berryman 1982c, Raffa 1988). Arrival rates to baited traps did<br />

not differ between stem phloem strips from root-infected <strong>or</strong> healthy trees, stems artificially infected with various fungi<br />

associated with Red Pine Decline, <strong>or</strong> volatiles associated with various intensities of stress (Table 2). When male I. pini were<br />

caged directly onto trees, however, the entry rates were 2.4 X higher on trees adjacent to the margin of dead trees than on<br />

trees farther into the stand. Pri<strong>or</strong> excavations had demonstrated high levels of root disease in the f<strong>or</strong>mer group (Klepzig et al.<br />

1991). Three conclusions can be drawn from these results: (1) beetles can detect changes in stem phloem physiology and<br />

Table 2.wComponents of I. pini colonization of P. resinosa affected by pri<strong>or</strong> infestation with root-colonizing insects and<br />

fungi<br />

Parameter Treatment Comparison Result<br />

a. Initial landing Host stem tissue Rootinfested vs. healthy trees ns<br />

Host stem tissue Fungal-infected vs. non-infected ns<br />

stems (Oi, On, Lt)<br />

Volatiles etOH, ct-pinene, EtOH & tx-pinene, ns<br />

vs. controls<br />

b. Host entry Caging onto whole trees Root infected vs. healthy trees Prefer diseased trees: 7.X<br />

c. Gallery f<strong>or</strong>mation Phloem strips Root weevil infected vs. healthy Prefer infested logs: 2.4X<br />

logs<br />

Amended phloem Nonpolai" extracts from constitu- Prefer healthy trees: 10X<br />

tive stem phloem of rootinfected<br />

vs. healthy trees<br />

Amended phloem Nonpolar extracts from reaction Prefer diseased trees: 3.0X<br />

stem phloem of root-infected<br />

vs. healthy trees<br />

Amended phloem Monoterpene contents present Prefer diseased trees: 3.5X<br />

in reaction stem phloem of<br />

root-infected vs. healthy trees<br />

d. Constitutive resin Resin flow Stem phloem of root-infected ns<br />

vs. healthy trees: 1.5m height<br />

Stem phloem of root-infected Higher in healthy trees: 4X<br />

vs. healthy trees: 10m height<br />

e. Reaction resin a-pinene Root diseased vs. healthy trees Higher in healthy trees: 1.3X<br />

216<br />

Data from Klepzig et al. (in prep.), Raffa and Smalley (in prep.).


chemistry associated with root colonization:; (2) there is no evidence that pre-landing behavi<strong>or</strong> is influenced by fact<strong>or</strong>s<br />

associated with root colonization; (3) the pattern of tree m<strong>or</strong>tality, in which trees near the "pocket" are m<strong>or</strong>e likely to be<br />

killed by I. pini than are trees distant from the margin of killed trees (Fig. 2), is not just a function of proximity to emerging<br />

brood. Even when the fact<strong>or</strong> of beetle dispersal is removed, higher entry rates are observed on neighb<strong>or</strong>ing trees.<br />

Beetle preference f<strong>or</strong> root-diseased over healthy trees persists under lab<strong>or</strong>at<strong>or</strong>y conditions. L pini excavated m<strong>or</strong>e<br />

extensive galleries in phloem strips of logs that had been artificially infested with H. radicis than in control logs. Likewise,<br />

extracts from tlhe stem phloem tissue of root-diseased and healthy trees affect L pini behavi<strong>or</strong> differently (Klepzig et al.<br />

1994c). When these extracts are applied to ground denatured phloem inc<strong>or</strong>p<strong>or</strong>ated into agar, male L pini show strong<br />

preferences. Both the extent of tunnelling and the location of beetles are influenced by the source of host extracts under twoway<br />

choice conditions. Interestingly, the beetles preferred extracts from the constitutive tissue of healthy over diseased trees.<br />

This suggests that low levels of monoterpenes function as feeding incitants. However, when trees are challenge inoculated<br />

with O. ips, the pattern is reversed. The reaction tissue of healthy trees is much m<strong>or</strong>e repellent than the reaction tissue of<br />

diseased trees (Table 2). The same results can be achieved by amending the media with synthetic monoterpene concentrations<br />

determined to be present in these various extracts.<br />

Once entered by beetles, root-infected trees are physiologically less able than uninfected trees to resist attack by/.<br />

pini and its associated fungus O. ips. Differences occur in both primary and induced resin (Table 3). Interestingly, we did not<br />

observe differences in primary resin flow rates between root-infected and healthy trees when the wounds were applied at<br />

1.5m height. However, flow rates at the base of the canopy yielded only 25% of the resin volume in root-diseased as in<br />

healthy trees. Thus, the earliest symptoms of root infection may arise relatively high up in the stem.<br />

Based on these results, we hypothesized the following sequence: (1) root- and lower stem- infecting beetles colonize<br />

selected trees and vect<strong>or</strong> Leptographium fungi into the root system; (2) root fungi progress through root grafts into healthy<br />

trees. These <strong>or</strong>ganisms do not kill mature red pines, but reduce their resistance to the stem-colonizing I. pini and its associated<br />

fungus O. ips; (3) colonization by <strong>this</strong> stem inhabiting complex is lethal to host trees; (4) killed trees provide a food base<br />

f<strong>or</strong> secondary root-feeding insects, which introduce additional inoculum into the epicenter.<br />

Table 3.wEnvironmentat fact<strong>or</strong>s affecting host suitability to insects and fungi associated with Red Pine Decline Disease<br />

Fact<strong>or</strong> Effect<br />

Reduced lights Increased lesion length follwoing inoculation with L. terebrantis: 2.5X<br />

Defoliation Increased lesion length following inoculation with O. ips: 1.2X<br />

Defoliation Decreased resin flow following mechanical wound: 0.45X<br />

Defoliation Preferential feeding by H. pales: 2.3X<br />

Nitrification Increased feeding by H. pales: 1.4X<br />

Nitrification Decreased development time by H. pales females: 0.89X<br />

Nitrification Decreased development time by H. pales males: 0.92X<br />

Nitrification Increased growth by H. pales females: 1.2X<br />

Nitrification Increased growth by H. pales males: 1.1X<br />

Nitrification incrased survival by H. pales: 2.1X<br />

Data from Hunt et al. (1993), Klepzig et al. (in 1994c.),Krause and Raffa (in prep.)<br />

217


Environmental Predisposition to Subc<strong>or</strong>tical Insects and Fungi Associated with Declining Red Pine Stands<br />

Although the above description may partially describe within-stand dynamics, it cannot explain the onset of decline.<br />

To address <strong>this</strong> question, we have conducted a variety of experiments on the effects of environmental stress on host suitability<br />

to insects and pathogens associated with Red Pine Decline. The results suggest that a relatively broad range of biotic and<br />

abiotic fact<strong>or</strong>s can improve host suitability to these <strong>or</strong>ganisms (Table 3). F<strong>or</strong> example, defoliation decreases resin flow,<br />

reduces tree ability to confine O. ips, and increases feeding preference to H. pales (Krause and Raffa 1994). Likewise,<br />

nitrification increases feeding, development, and survival by H. pales (Hunt et al. 1993). Reduced light availability also<br />

decreases host ability to respond to L. terebrantis (Klepzig et al. 1994c). Both defoliation and reduced light availability<br />

reduce the overall pool of photosynthate f<strong>or</strong> carbon-based defenses such as terpenes and phenolics (Wright et al. 1979,<br />

Ericsson et al. 1985, L<strong>or</strong>io and Summers 1986, Miller and Berryman 1986, Peet and Christensen 1987, Waring 1987, Dunn et<br />

al. 1990, Reich et al. 1992, Krause et al. 1993, Krause and Raffa 1994). These stresses can themselves be initiated by biotic<br />

agents. F<strong>or</strong> example, outbreaks by insect (<strong>or</strong> microbial) defoliat<strong>or</strong>s are often followed by m<strong>or</strong>tality due to bark beetles<br />

(Wright et al. 1984, Paine and Baker 1993), densely crowded stands m<strong>or</strong>e commonly experience bark beetle outbreaks<br />

(Nebeker et al. 1993), and suppressed trees are less able to resist attack (Safranyik et al. 1975, Raffa and Berryman 1982b,<br />

Mitchell et al. 1983, Waring and Pitman 1985, Paine and Baker 1993, Preisler and Mitchell 1993).<br />

Our current understanding of the biotic and abiotic interactions contributing to Red Pine Decline is depicted in Figure<br />

4. Each of the maj<strong>or</strong> assumptions underlying <strong>this</strong> model has been supp<strong>or</strong>ted, although some of the mechanisms are only<br />

po<strong>or</strong>ly understood. In particular, our knowledge of the relative imp<strong>or</strong>tance, impact, and tolerance ranges of the various<br />

inciting fact<strong>or</strong>s lags behind our knowledge of within-stand spread. Our current understanding is now sufficient, however, to<br />

test putative natural and anthropogenic stress agents on specific tree physiological, insect behavi<strong>or</strong>al and developmental, and<br />

microbial developmental parameters, and to project these impacts to the stand level. This model can also serve as a framew<strong>or</strong>k<br />

f<strong>or</strong> expanding our basic understanding of interactions among feeding guilds, the interaction of constitutive and inducible<br />

allelochemicals in host plant resistance, and microbial mediation of plant-insect interactions.<br />

.,__,_._ _:_ • ..., _,_...,_...,, :- . :. ..<br />

..... g g g ..... ..... .<br />

Figure 4._Sequence of events in the infection, decline, and m<strong>or</strong>tality of plantation red pines.<br />

218


Implications to Natural Resource Management and F<strong>or</strong>est Ecosystem Health<br />

The dynamics of insect and microbial colonization also suggests some specific management tactics that can prevent<br />

the onset and spread of Red Pine Decline. F<strong>or</strong> example, mixed-block rather than unif<strong>or</strong>m-species planting would reduce the<br />

transmission of Leptographium fungi through root grafts. This would contain root-diseased trees within relatively small<br />

groups rather than entire stands. Where economic incentives f<strong>or</strong> red pine make large-scale planting of"alternate species<br />

impractical, planting buffer strips of alternate species might serve the same purpose. Pinus strobus would probably be a<br />

better candidate than Pinus banksiana, because the f<strong>or</strong>mer is less preferred by the root weevils in our region (Hunt et al.<br />

1993). Judicious thinning would also reduce stress and thereby reduce stand susceptibility. However, the timing is quite<br />

imp<strong>or</strong>tant. Open light penetration can fav<strong>or</strong> site suitability to root weevils by warming and drying the soil (Wilson and<br />

Millers 1983). M<strong>or</strong>eover, thinning should be accompanied by stump removal to avoid leaving a food base f<strong>or</strong> the root- and<br />

lower-stem feeding species (Witcosky and Hansen 1985, Witcosky et al. 1986). Our model also suggests that severing root<br />

grafts in advance (@ 15m) of killed trees could halt the transmission of Leptographium. We have begun tests on <strong>this</strong> idea in<br />

Wisconsin. We further recommend sanitation removal of dead and living trees (including stumps) within the radius of root<br />

severing.<br />

Although the insects and fungi described here are typically viewed as f<strong>or</strong>est pests, these moderately pathogenic<br />

species might serve as useful bio-indicat<strong>or</strong>s. Reproductive success by these <strong>or</strong>ganisms is contingent on their ability to detect<br />

trees that are sufficiently stressed to have impaired defensive capability, but not so severely stressed as to be available to<br />

saprophytic competit<strong>or</strong>s. As such, the population densities and behavi<strong>or</strong> of moderately virulent endoparasites may be<br />

indicative of reduced ecosystem vig<strong>or</strong> associated with anthropogenic disturbances <strong>or</strong> injudicious f<strong>or</strong>estry practices. Our<br />

monit<strong>or</strong>ing data and bioassay show <strong>this</strong> idea is feasible f<strong>or</strong> at least one f<strong>or</strong>m of f<strong>or</strong>est decline. That is, populations of some<br />

root- and lower stem-feeding insects were higher in declining than in asymptomatic stands. M<strong>or</strong>eover, different species vary<br />

in how their densities relate to the progression of decline and in how they respond to tree stress physiology. We propose that<br />

the most useful bioindicat<strong>or</strong> candidates might be those that show increased numbers <strong>or</strong> behavi<strong>or</strong>al alteration bef<strong>or</strong>e any<br />

symptoms become apparent, but are not by themselves maj<strong>or</strong> outbreak pests.<br />

SUMMARY<br />

Interactions between feeding guilds that affect different plant parts can strongly affect the dynamics of f<strong>or</strong>est declines.<br />

Environmental stresses can mediate these outcomes, especially against highly generalized plant defense systems. Such<br />

interactions among insects and fungi are primarily responsible f<strong>or</strong> Red Pine Decline, a syndrome affecting mature plantation<br />

trees in the Great Lakes region of N<strong>or</strong>th America. Root- and lower stem-colonizing insect-fungal complexes reduce tree<br />

resistance against stem-feeding bark beetle-fungal complexes. Between-stand transmission of moderately pathogenic<br />

Ophiostomatales fungi occurs by insect vect<strong>or</strong>ing, but within-stand progression also occurs through extensive root grafts in<br />

plantations. Subterranean scolytids and curculionids show a high degree of niche partitioning based on host tissue and<br />

condition, chemical ecology, microbial associates, and spatial and temp<strong>or</strong>al distributions.<br />

Healthy trees can resist colonization by rapidly accumulating terpenes and phenolics to concentrations that repel <strong>or</strong><br />

kill I. pini and inhibit O. ips sp<strong>or</strong>e germination and mycelial growth. However, tree constitutive and inducible defenses are<br />

compromised by root infection. Behavi<strong>or</strong>al assays are generally the most sensitive to the early stages of root injury. Among<br />

these, post landing assays of I. pini acceptance behavi<strong>or</strong> are most sensitive, especially those in which beetles are exposed to<br />

host reaction tissue f<strong>or</strong>med in response to challenge inoculation with fungi. Host susceptibility to the root colonizing agents<br />

can be increased by environmental stresses such as reduced light, defoliation, and nitrification.<br />

An understanding of the complex interactions underlying f<strong>or</strong>est declines can improve our ability to relate various<br />

syndromes to management practices, anthropogenic stress, and natural succession. These systems can also serve as useful<br />

models to improve our overall understanding of plant-insect and insect-fungal interactions. The underlying mechanisms also<br />

suggest some specific tactics f<strong>or</strong> preventing <strong>or</strong> ameli<strong>or</strong>ating Red Pine Decline. Marginally pathogenic agents such as the<br />

<strong>or</strong>ganisms described here may also serve as useful indicat<strong>or</strong>s of f<strong>or</strong>est ecosystem health. Because their existence is contingent<br />

on detecting trees under sufficient stress to fully resist attack but not yet available to saprophytic competit<strong>or</strong>s, the<br />

population densities, behavi<strong>or</strong>, and developmental success of these insects and fungi n-fight be good indicat<strong>or</strong>s of incipient<br />

stress at the stand <strong>or</strong> tree level.<br />

219


ACKNOWLEDGMENTS<br />

This w<strong>or</strong>k was supp<strong>or</strong>ted by <strong>USDA</strong> CSRS Competitive Grant 90-3725P-5581, the University of Wisconsin Graduate<br />

School, the University of Wisconsin College of Agricultural and Life Sciences, Mclntire Stennis, and Hatch Regional Project<br />

W-187. We would also like to thank our colleagues L. Rieske, S. Krause, and K. Kleiner )'<strong>or</strong> input on the ideas discussed in<br />

<strong>this</strong> paper.<br />

REFERENCES CITED<br />

BERRYMAN, A.A. 1972. Resistance of conifers to invasion by bark beetle-fungus associations. Bio Science 22: 598-602.<br />

CATES, R.G. and ALEXANDER, H. 1982. Host resistance and susceptibility, p. 212-263. In Mitton, J.B. and Sturgeon,<br />

K.B., eds. Bark Beetles in N<strong>or</strong>th American Conifers: A System f<strong>or</strong> the Study of Evolutionary Biology. University<br />

Texas Press, Austin.<br />

CHRISTENSEN, E. and ERICSSON, A. 1986. Starch reserves in Picea abies in relation to defense reaction against a bark<br />

beetle transmitted blue-stain fungus, Ceratocuptispolonica. Can. J. F<strong>or</strong>. Res. 16: 78-83.<br />

CHRISTENSEN, E. and HORNTVEDT, R. 1983. Combined lps Ceratocystis attack on N<strong>or</strong>way spruce and defensive<br />

mechanisms of the trees. Zeit. angew. Entomol. 96:110-118.<br />

DUNN, J.E, POTTER, D.A., and KIMMERER, T.W. 1990. Carbohydrate reserves, radial growth, and mechanisms of<br />

resistance of oak trees to phloem-b<strong>or</strong>ing insects. Oecologia 83: 458-468.<br />

ELKINTON, J.W. and WOOD, D.L. 1980. Feeding and b<strong>or</strong>ing behavi<strong>or</strong> of the bark beetle lps paraconfusus (Coleoptera<br />

Scolytidae) on the bark of a host and non-host tree species. Can. Entomol. 112: 797-809.<br />

ERICSSON, A., HELLQUIST, C., LANGSTROM, B., LARSSON, S., and TENNOW, O. 1985. Effects on growth of<br />

simulated and induced shoot pruning by Tomicus piniperda as related to carbohydrate and nitrogen dynamics in Scots<br />

pine. J. Applied Ecol. 22: 105-124.<br />

HODGES, J.D., ELAM, W.W., WATSON, W.E, and NEBEKER, T.E. 1979. Ole<strong>or</strong>esin characteristics and susceptibility f<strong>or</strong><br />

four southern pines to southern pine beetle attacks. Can. Entomol. 11: 889-896.<br />

HOUSTON, D.R. 1981. Stress triggered tree diseases_the diebacks and declines. NE-INF-41-81. Radn<strong>or</strong>, PA: U.S.<br />

Department of Agriculture, F<strong>or</strong>est Service, N<strong>or</strong>theastern F<strong>or</strong>est Experiment <strong>Station</strong>: 4-7.<br />

HUETTL, R.E, FINK, S., LUTZ, H-J., POTH, M., and WISMEWSKI, J. 1990. F<strong>or</strong>est decline, nutrient supply and diagnostic<br />

fertilization in Southwestern Germany and in Southern Calif<strong>or</strong>nia. F<strong>or</strong>. Ecol. and Manage. 30: 341-350.<br />

HUNT, D.W.A., LINTEREUR, G., SALOM, S., and RAFFA, K.E 1993. Perf<strong>or</strong>mance and preference of Hylobius radicis<br />

Buchanan, and H. pales (Herbst) (Coleoptera:Curculionidae)on various Pinus species. Can. Entomol. 125:1003-<br />

1010.<br />

KIMMERER, T.W. and KOZLOWSKI, T.T. 1982. Ethylene, ethane, acetyl aldehyde, and ethanol production by plants<br />

under stress. Plant Physiol. 69: 840-847.<br />

KLEIN, R.M. and PERKINS, T.D. 1988. Primary and secondary causes and consequences of contemp<strong>or</strong>ary f<strong>or</strong>est decline.<br />

The Bot. Rev. 54: 1-43.<br />

KLEPZIG, K.D., RAFFA, K.E, and SMALLEY, E.B. 1991. Association of insect-fungal complexes with Red Pine Decline<br />

in Wisconsin. F<strong>or</strong>. Sci. 37: 1119-1139.<br />

220


KLEPZIG, K.D., RAFFA, K.F., and SMALLEY, E.B. 1994a. Reaction of red pine to inoculation with isolates of the fungi<br />

associated with Red Pine Decline Disease. In prep_<br />

KLEPZIG, K.D., RAFFA, K.F., and SMALLEY, E.B. 1994b. Dendroctonus valens and Hylastes p<strong>or</strong>culus: vect<strong>or</strong>s of<br />

pathogenic fungi (Ophiostomatales) associated with Red Pine Decline Disease. Subm. to Environ. Entomol.<br />

KLEPZlG, K.D., KRUGER, E.L., SMALLEY, E.B., and RAFFA, K.F. 1994c. Monoterpenes and phenolics associated with<br />

the response of healthy and diseased red pine to bark beetle vect<strong>or</strong>ed fungi. In prep.<br />

KLEPZlG, K.D., RIBA, J., SMALLEY, E.B., and RAFFA, K.F. 1994d. Effects of red pine phenolics and monoterpenes on a<br />

subc<strong>or</strong>tical insect - fungal complex associated with a decline disease. In prep.<br />

KRAUSE, S. and RAFFA, K.F. 1994. Effects of long-term defoliation on insect herbiv<strong>or</strong>es of red pine. In prep.<br />

KRAUSE, S.C., RAFFA, K.F., and WAGNER, M.R. 1993. Tree response to stress: a role in sawfly outbreaks? p. 211-227.<br />

In Wagner, M.R. and Raffa, K.F., eds. Sawfly Life Hist<strong>or</strong>y Adaptations to Woody Plants. Academic Press, New<br />

Y<strong>or</strong>k.<br />

LIEUTIER, F. and BERRYMAN, A.A. 1988. Elicitation of defensive reactions in conifers, p. 313-320. In Mattson, W.J.,<br />

Levieux, J., and Bernard-Dagan, C., eds. Mechanisms of Woody Plant Defenses Against Insects: Search f<strong>or</strong> Pattern.<br />

Springer-Verlag, New Y<strong>or</strong>k.<br />

LORIO, P.L., JR. 1986. Growth-differentiation balance: a basis f<strong>or</strong> understanding southern pine beetle-tree interactions.<br />

F<strong>or</strong>. Ecol. & Manage. 14: 259-273.<br />

LORIO, RL., JR. 1993. Environmental stress and whole-tree physiology, p. 81-101. In Schowalter, R.D. and Filip, G.M.,<br />

eds. Beetle-Pathogen Interactions in Conifer F<strong>or</strong>ests. Academic Press, New Y<strong>or</strong>k.<br />

LORIO, RL., JR. and SUMMERS, R.A. 1986. Evidence of competition f<strong>or</strong> photosynthates between growth processes and<br />

ole<strong>or</strong>esin synthesis in Pinus taeda. Tree Physiol. 2: 301-306.<br />

MANION, RD. 1991. Decline diseases of complex biotic and abiotic <strong>or</strong>igin, p. 324-339. In Manion, RD., ed. Tree Disease<br />

Concepts. Prentice Hall, Englewood Cliffs, NJ.<br />

MILLER, R.H. and BERRYMAN, A.A. 1986. Carbohydrate allocation and mountain pine beetle attack in girdled lodgepole<br />

pines. Can. J. F<strong>or</strong>. Res. 16: 1036-1040.<br />

MILLER, R.H., WHITNEY, H.S., and BERRYMAN, A.A. 1986. Effects of induced translocation stress and bark beetle<br />

attack (Dendroctonus ponderosae) on heat pulse velocity and the dynamic wound response of lodgepole pine (Pinus<br />

cont<strong>or</strong>ta var. latifolia). Can. J. Bot. 64: 2669-2674.<br />

MITCHELL, R.G., WARING, R.H., and PITMAN, G.B. 1983. Thinning lodgepole pine increases tree vig<strong>or</strong> and resistance<br />

to mountain pine beetle. F<strong>or</strong>. Sci. 29: 204-211.<br />

MUELLER-DOMBOIS, D. 1988. F<strong>or</strong>est decline and die backmA global ecological problem. Trends in Ecol. and Evol. 3:<br />

310-312.<br />

NEBEKER, T.E., HODGES, J.D., and BLANCHE, C.A. 1993. Host response to bark beetle and pathogen colonization, p.<br />

157-178. In Schowalter, R.D. and Filip, G.M., eds. Beetle-Pathogen Interactions in Conifer F<strong>or</strong>ests. Academic<br />

Press, New Y<strong>or</strong>k.<br />

PAINE, T.D., and BAKER, F.A. 1993. Abiotic and biotic predisposition, p. 61-79. In Schowalter, R.D. and Filip, G.M., eds.<br />

Beetle-Pathogen Interactions in Conifer F<strong>or</strong>ests. Academic Press, New Y<strong>or</strong>k.<br />

221


PAINE, T.D. and STEPHEN, F.M.. 1987a. The relationship of tree height and crown class to the induced plant defenses of<br />

loblolly pine. Can. J. Bot. 65: 2090-2092.<br />

PAINE, T.D. and STEPHEN, F.M. 1987b. Influence of tree stresses and site quality on the induced defensive system of<br />

loblolly pine. Can. J. F<strong>or</strong>. Res. 17: 569-571.<br />

PEET, R.K. and CHRISTENSEN, N.L. 1987. Competition and tree death. BioScience 37::587-595.<br />

PREISLER, H.K. and MITCHELL, R.G. 1993. Colonization patterns of the mountain pine beetle in thinned and unthinned<br />

lodgepole pine stands. F<strong>or</strong>. Sci. 39 528-545.<br />

RAFFA, K.F. 1988. Host <strong>or</strong>ientation behavi<strong>or</strong> of Dendroctonusponderosae: Integration of token stimuli and host defenses,<br />

p. 369-390. In Mattson, W.J., Levieux, J., and Bernard-Dagan, C., eds. Mechanisms of Woody Plant Defenses<br />

Against Insects: Search f<strong>or</strong> Pattern. Springer-Verlag, New Y<strong>or</strong>k.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1982a. Physiological differences between lodgepole pines resistant and susceptible<br />

to the mountain pine beetle and associated micro<strong>or</strong>ganisms. Environ. Entomol. 11: 486-492.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1982b. Accumulation of monoterpenes and associated volatiles following fungal<br />

inoculation of grand fir with a fungus transmitted by the fir engraver, Scolytus ventralis (Coleoptera: Scolytidae).<br />

Can. Entomol. 114:797-810.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1982c. Gu_tat<strong>or</strong>y cues in the <strong>or</strong>ientation of Dendroctonus ponderosae<br />

(Coleoptera:Scolytidae) to host trees. Can. Entomol. 114: 97-104.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1983a. Physiological aspects of lodgepole pine wound responses to a fungal symbiont<br />

of the mountain pine beetle. Can. Entomol. ! 15: 723-734.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1983b. The role of host plant resistance in the colonization behavi<strong>or</strong> and ecology of<br />

bark beetles. Ecol. Monogr. 53: 27-49.<br />

RAFFA, K.F. and BERRYMAN, A.A. 1987. Interacting selective pressures in conifer-bark beetle systems" A basis f<strong>or</strong><br />

reciprocal adaptations? Am. Nat. 129: 234-262.<br />

RAFFA, K.F. and HUNT, D.W. 1989. Microsite and interspecific interactions affecting emergence of root-infesting pine<br />

weevils (Coleoptera: Curculionidae) in Wisconsin. Ann. Entomoi. Soc. Am. 82: 438-445.<br />

RAFFA, K.F. and KLEPZIG, K.D. 1992. Tree defense mechanisms against insect-vect<strong>or</strong>ed fungi, p. 354-390. In<br />

Blanchette, R.A. and Biggs, A.R., eds. Defense Mechanisms of Woody Plants against Fungi. Springer-Verlag, New<br />

Y<strong>or</strong>k.<br />

RAFFA, K.F. and SMALLEY, E.B. 1988a. Host resistance to invasion by lower stem and root infesting insects of pine:<br />

Response to controlled inoculations with the fungal associate Leptographium terebrantis. Can. J. F<strong>or</strong>. Res. 18" 675-<br />

681.<br />

RAFFA, K.F. and SMALLEY, E.B. 1988b. Response of red and jack pines to inoculation with microbial associates of the<br />

pine engraver, lps pini. Can. J. F<strong>or</strong>. Res. 18:581-586.<br />

RAFFA, K.F. and SMALLEY, E.B. 1988c. Seasonal and long-term responses of host trees to microbial associates of the<br />

pine engraver, Ips pini. Can. J. F<strong>or</strong>. Res. 18: 1624-1634.<br />

RAFFA, K.F. and SMALLEY, E.B. 1994. Induction of red pine subc<strong>or</strong>tical monoterpenes to insecticidal levels by fungal<br />

associates of bark beetles. Subm. to Oecologia.<br />

222


RAFFA, K.F., PHILLIPS, T., and SALOM, S. 1993. Strategies and mechanisms of host colonization by bark beetles, p. 103-<br />

128. In Schowalter, T.O. and Filip, G., eds. Beetle-Pathogen Interactions in Conifer F<strong>or</strong>ests. Academic Press, New<br />

Y<strong>or</strong>k.<br />

REICH, RB., WALTERS, M.B., KRAUSE, S.C., VANDERKLEIN, D.W., RAFFA, K.F., and TIBONE, T. 1992. Gas<br />

exchange and growth of red pine seedlings and trees one year after defoliation. Trees - Structure and Function. 7: 67-<br />

77.<br />

RIESKE, L.K. and RAFFA, K.F. 1990. Dispersal patterns and mark-recapture estimates of two pine root weevil species<br />

Hylobius pales and Pachylobius piciv<strong>or</strong>us (Coleoptera: Curculionidae), in Christmas tree plantations. Environ.<br />

Entomol. 19: 1829-1836.<br />

RIESKE, L.K. and RAFFA, K.F. 1991. Effect of varying ethanol and turpentine levels on attraction of two pine root weevil<br />

species Hylobius pales and Pachylobius piciv<strong>or</strong>us (Coleoptera: Curculionidae). Environ. Entomol. 20: 48-52.<br />

SAFRANYIK, L., SHRIMPTON, D.M., and WHITNEY, H.S. 1975. An interpretation of the interaction between lodgepole<br />

pine, the mountain pine beetle and its associated blue stain fungi in western Canada, p. 406-428. In Baumgartner,<br />

D.M., ed. Management of Lodgepole Pine Ecosystems. WSU Coop. Ext. Serv.<br />

SAXE, H. 1993. Triggering and predisposing fact<strong>or</strong>s in the "Red" decline syndrome of N<strong>or</strong>way spruce (Picea abies). Trees<br />

8: 39-48.<br />

WARING, R.H. 1987. Characteristics of trees predisposed to die. BioScience 37: 569-574.<br />

WARING, R.H. and PITMAN, G.B. 1985. Modifying lodgepole pine stands to change susceptibility to mountain pine beetle<br />

attacks. Ecology 66: 889-897.<br />

WILSON, L.F. and MILLERS, I. 1983. Pine root collar weevil: Its ecology and management. Tech. Bull. 1675. Washington,<br />

DC: U.S. Department of Agriculture, F<strong>or</strong>est Service.<br />

WITCOSKY, J.J. and HANSEN, E.M. 1985. Root-colonizing insects associated with Douglas-fir in various stages of<br />

decline due to black-stain root disease. Phytopathol. 75: 399-402.<br />

WlTCOSKY, J.J., SCHOWALTER, T.D., and HANSEN, E.M. 1986. The influence of time of precommercial thinning on<br />

the colonization success of Douglas-fir by three species of root-colonizing insects. Can. J. F<strong>or</strong>. Res. 16: 745-749.<br />

WOOD, D.L. 1972. Selection and colonization of ponderosa pine by bark beetles. Symposium of the Royal Entomol. Soc.<br />

London. 6:110-117.<br />

WOODCOCK, H., PATTERSON, W.A. III, and DAVIES, EM., JR. 1993. The relationship between site fact<strong>or</strong>s and white<br />

ask (Fraximus americana L.) decline in Massachusetts. F<strong>or</strong>. Ecol. & Manag. 60:271-290.<br />

WRIGHT, L.E., BERRYMAN, A.A., and GURUSIDDIAH, S. 1979. Host resistance to the fir engraver beetle, Scolytus<br />

ventralis (Coleoptera:Scolytidae). 4. Effect of defoliation on wound monoterpenes and inner bark carbohydrate<br />

concentrations. Can. Entomol. 111: 1255-1261.<br />

WRIGHT, L.A., BERRYMAN, A.A., and WICKMAN, B.E. 1984. Abundance of the fir engraver, Scolytus ventralis, and<br />

the Douglas-fir beetle, Dendroctonus pseudotsugae, following tree defoliation by the Douglas-fir tussock moth,<br />

Orgyria pseudotsugata. Can. Entomol. 116: 293-305.<br />

223


DOUGLAS-FIR AND WESTERN LARCH DEFENSIVE REACTIONS TO<br />

LEPTOGRAPHIUM ABIETINUM AND OPHIOSTOMA PSEUDOTSUGAE<br />

DARRELL W. ROSS 1and HALVOR SOLHEIM 2<br />

_Department of F<strong>or</strong>est Science, Oregon State University, C<strong>or</strong>vallis, Oregon, 97331-7501, USA<br />

2N<strong>or</strong>wegian F<strong>or</strong>est <strong>Research</strong> Institute, Section of F<strong>or</strong>est Ecology, Division of F<strong>or</strong>est Pathology, N-1423 As-NLH, N<strong>or</strong>way<br />

INTRODUCTION<br />

The Douglas-fir beetle, Dendroctonus pseudotsugae Hopkins, is one of the most imp<strong>or</strong>tant insects associated with<br />

Douglas-fir, Psuedotsugae menziesii (Mirb.) Franco, in western N<strong>or</strong>th America (Furniss and Carolin 1977). At high population<br />

densities, the beetle is capable of causing significant tree m<strong>or</strong>tality and impacting f<strong>or</strong>est resource values (C<strong>or</strong>nelius 1955,<br />

Rudinsky 1966, Johnson and Belluschi 1969, Furniss and Orr 1978, Furniss et al. 1979). The Douglas-fir beetle usually<br />

breeds in recently dead <strong>or</strong> live Douglas-fir trees. The only other tree species in which the beetle has been rep<strong>or</strong>ted to successfully<br />

breed is western larch (Ross 1967). However, the beetle is apparently able to breed only in dead larch (Furniss et al.<br />

1981). Live larch may be attacked, but brood production from live trees has never been observed. Reed et al. (1986)<br />

compared various chemical and physical properties of Douglas-fir and western larch to identify possible reasons that brood<br />

fail to survive in live larch trees. Compared with Douglas-fir, larch had a higher content of 3-carene, thinner phloem, higher<br />

phloem moisture content, larger diameter vertical resin ducts, and lower ole<strong>or</strong>esin exudation pressure. The auth<strong>or</strong>s speculated<br />

that the high content of 3-carene in live larch may be responsible f<strong>or</strong> the failure of Douglas-fir beetle brood in those<br />

trees.<br />

One <strong>or</strong> m<strong>or</strong>e species of ophiostomatoid fungi are invariably associated with each species of conifer-infesting bark<br />

beetle (Whitney 1982, Harrington 1988). Although the ecological relationships among the fungi and bark beetles are not<br />

completely understood in many cases, there is evidence that some of these fungi help the beetles to overcome the natural<br />

defenses of mass attacked trees. Conifer response to invasion by bark beetles and their associated fungi involves two<br />

different systems: (1) the flow of pref<strong>or</strong>med resin from severed resin ducts and (2) the induced wound response <strong>or</strong> resistant<br />

reaction (Reid et al. 1967, Berryman 1972, Christiansen and H<strong>or</strong>ntvedt 1983). The dimensions of the induced response in<br />

host tree tissues are generally c<strong>or</strong>related to the growth of the fungi <strong>or</strong> the extent of mechanical damage associated with beetle<br />

tunneling (Ross et al. 1992, Lieutier 1993, Solheim 1993). If the fungi associated with the Douglas-fir beetle are involved in<br />

overcoming tree resistance, then a possible reason f<strong>or</strong> the failure of brood development in larch may be the inability of the<br />

fungi to survive and grow in larch. A significant variation in the pathogenicity of the fungi to Douglas-fir and western larch<br />

should he reflected by differences in the induced response following artificial inoculations.<br />

Despite a considerable amount of past research related to the biology and management of the Douglas-fir beetle, little<br />

is known about its funsal associates. The two species of ophiostomatoid fungi most commonly associated with the Douglasfir<br />

beetle are Ophiostoma pseudotsugae (Rumb.) yon Arx and Leptographium abietinum (Peck) Wingf. (Rumbold 1936,<br />

Harrington 1988, Solheim unpubl.). There are no published rep<strong>or</strong>ts on the pathogenicity of these fungi <strong>or</strong> their ecological<br />

relationships to the beetle. In 1993, we installed a study to evaluate the pathogenicity of these fungi to Douglas-fir. The final<br />

results of that study will be published elsewhere in the near future. At the same time, we installed a small test to assess the<br />

response of Douglas-fir and western larch, Larix occidentalis Nutt., to artificial inoculations with these fungi. The objective<br />

of our study was to measure the length of the induced response in the phloem of Douglas-fir and western larch following<br />

artificial inoculation with O. pseudotsugae and L. abietinum. This paper rep<strong>or</strong>ts the results of <strong>this</strong> test.<br />

Mattson, W.J., Niemela, E, and Rouse, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108. i<br />

224


METHODS<br />

The study was installed in a mixed conifer stand (elevation - 1460 m) in n<strong>or</strong>theastern Oregon. On June 29, 1993,<br />

seven Douglas-fir and seven western larch were selected from an area of less than 1 ha. The trees were chosen based on<br />

similarities in size and lack of any visible injuries <strong>or</strong> disease. Three inoculation sites were established equidistant around the<br />

circumference of each tree at dbh (1.4 m height). The three treatments that were randomly assigned to the inoculation sites<br />

were: (1) sterile wound, (2) L. abietinum, and (3) O. pseudotsugae. At each inoculation site, a 1.2-cm diameter hole was<br />

drilled to the sapwood surface with a sterile bit. F<strong>or</strong> the sterile wound, a cotton plug was placed in the hole that was then<br />

covered with duct tape. F<strong>or</strong> the fungal inoculations, a small agar plug containing the appropriate fungus was placed in the<br />

hole that was then similarly covered. The fungi used in the study had been isolated from Douglas-fir beetles collected in<br />

British Columbia, Canada (Solheim unpubl.).<br />

On August 25, 1993, the outer bark covering each inoculation site was removed and the length of the necrotic lesion<br />

on the phloem surface was measured.<br />

The mean dbh f<strong>or</strong> trees of each species were compared by a t-test. Lesion length was subjected to analysis of<br />

variance f<strong>or</strong> a split-plot design with tree species representing whole plots and fungal inoculation representing subplots.<br />

RESULTS<br />

There was no significant difference (P < 0.19) in dbh between Douglas-fir (mean = 51.0 cm) and western larch (mean<br />

= 48.8 cm). The tree species x treatment interaction was significant (F = 3.29; P < 0.05) f<strong>or</strong> lesion length. F<strong>or</strong> both tree<br />

species, inoculation with either fungus resulted in a larger lesion than that produced by a sterile wound (Fig. 1). The length<br />

of the lesions induced by a sterile wound <strong>or</strong> inoculation with O. pseudotsugae did not differ between tree species, but the<br />

lesion induced by inoculation with L. abietinum was significantly sh<strong>or</strong>ter in western larch than in Douglas-fir (Fig. 1).<br />

DISCUSSION<br />

There were no obvious differences in the general appearance of lesions (i.e., intensity of resin accumulation, radial<br />

growth, etc.) between tree species, so lesion length should provide a relatively accurate indication of fungal growth within the<br />

phloem of the host trees. The smaller lesions produced by western larch in response to L. abietinum may reflect a m<strong>or</strong>e rapid<br />

and effective defensive reaction to the fungus. These results suggest that L. abietinum is not as well adapted to utilize<br />

western larch as a host compared with Douglas-fir. If L. abietinum is critical to the survival of Douglas-fir beetle brood in<br />

live host trees, then <strong>this</strong> may help to explain the complete m<strong>or</strong>tality of brood in western larch. In contrast, there was no<br />

difference in the induced response to O. pseudotsugae between the two tree species. Our results must be interpreted cautiously<br />

since we have only measured lesion length in phloem tissue. Lesion length is not always directly c<strong>or</strong>related to<br />

pathogenicity (Solheim 1988). Fungal growth and tree response may be quite different in the xylem compared with the<br />

phloem. However, the difference in response to fungal inoculations that we observed suggests that further research on the<br />

host tree-fungus interactions in <strong>this</strong> system would be w<strong>or</strong>thwhile. The significance of these results will only become apparent<br />

when we learn m<strong>or</strong>e about the role that these fungi play in the population dynamics of the Douglas-fir beetle.<br />

SUMMARY<br />

There was no difference in lesion length at the phloem surface between Douglas-fir and western larch following<br />

sterile wounds <strong>or</strong> inoculation with O. pseudotsugae. However, lesion length following inoculation with L. abietinurn was<br />

significantly sh<strong>or</strong>ter in western larch compared with Douglas-fir. These results may help to explain the lack of Douglas-fir<br />

beetle brood survival in western larch.<br />

225


5O<br />

40-<br />

Douglas-fir a<br />

a<br />

Western larch a<br />

E<br />

E b<br />

30-<br />

J:E<br />

E<br />

"-- C<br />

E 20o<br />

(_<br />

_J<br />

10-<br />

0<br />

Wound L. ab/et/num O. pseudotsugae<br />

Treatment<br />

Figure 1.--Mean length of the induced lesion at the phloem surface following sterile wounding <strong>or</strong> fungal inoculation in<br />

Douglas-fir and western larch (n=7). Bars with the same letter are not significantly different (P = 0.05).<br />

LITERATURE CITED<br />

BERRYMAN, A.A. 1972. Resistance of conifers to invasion of bark beetle-fungus associations. BioScience 22: 598-602.<br />

CHRISTIANSEN, E. and HORNTVEDT, R. 1983. Combined Ips/Ceratocystis attack on N<strong>or</strong>way spruce, and defensive<br />

mechanisms of the trees. Z. Ang. Entomol. 96:110-118.<br />

CORNELIUS, R.O. 1955. How f<strong>or</strong>est pests upset management plans in the Douglas-fir region. J. F<strong>or</strong>. 53:71 I-713.<br />

FURNISS, M.M. and ORR, P.W. 1978. Douglas-fir beetle. F<strong>or</strong>. Insect & Dis. Leafl. 5. Washington, DC" U.S. Department<br />

of Agriculture, F<strong>or</strong>est Service.<br />

FURNISS, M.M., MCGREGOR, M.D., FOILES, M.W., and PARTRIDGE, A.D. 1979. Chronology and characteristics of a<br />

Douglas-fir beetle outbreak in n<strong>or</strong>thern Idaho. Gen. Tech. Rep. INT-59. Ogden UT: U.S. Department of Agriculture,<br />

F<strong>or</strong>est Service.<br />

226<br />

: ii


FURNISS, M.M., LIVINGSTON, R.L., and MCGREGOR, M.D. 1981. Development of a stand susceptibility classification<br />

f<strong>or</strong> Douglas-fir beetle, p. 115-128. In Hedden, R.L., Barras, S.J., and Coster, J.E., tech. co<strong>or</strong>ds. Hazard-Rating<br />

Systems in F<strong>or</strong>est Insect Pest Management: Symposium Proceedings. Gen. Tech. Rep. WO-27. Washington, DC:<br />

U.S. Department of Agriculture, F<strong>or</strong>est Service.<br />

FURNISS, R.L. and CAROLIN, V.M. 1977. Western f<strong>or</strong>est insects. Misc. Publ. 1339. Washington, DC: U.S. Department<br />

of Agriculture, F<strong>or</strong>est Service.<br />

HARRINGTON, T.C. 1988. Leptograhium species, their distributions, hosts and insect vect<strong>or</strong>s, p. 1-40. In Harrington, T.C.<br />

and F.W. Cobb, Jr., eds. Leptographium Root Diseases on Conifers. American Phytopathological Society Press, St.<br />

Paul, MN.<br />

JOHNSON, N.E. and BELLUSCHI, RG. 1969. ttost-finding behavi<strong>or</strong> of the Douglas-fir beetle. J. F<strong>or</strong>. 67: 290-295.<br />

LIEUTIER, E 1993. Induced defence reaction of conifers to bark beetles and their associated Ophiostoma species, p. 225-<br />

234. In Wingfield, M.J., Seifert, K.A., and Webber, J.E, eds. Ceratocystis and Ophiostoma Taxonomy, Ecology, and<br />

Pathogenicity. American Phytopathological Society Press, St. Paul, MN.<br />

REED, A.N., HANOVER, J.W., and FURNISS, M.M. 1986. Douglas-fir and western larch: chemical and physical properties<br />

in relation to Douglas-fir bark beetle attack. Tree Physiol. 1: 277-287.<br />

REID, R.W., WHITNEY, H.S., and WATSON, J.A. 1967. Reactions of lodgepole pine to attack by Dendroctonus<br />

ponderosae Hopkins and blue stain fungi. Can. J. Bot. 45:1115-1126.<br />

ROSS, D.A. 1967. Wood- and bark-feeding Coleoptera of felled western larch in British Columbia. J. Entomol. Soc. Brit ....<br />

Columbia 64: 23-24.<br />

ROSS, D.W., FENN, R, and STEPHEN, F.M. 1992. Growth of southern pine beetle associated fungi in relation to the<br />

induced wound response in loblolly pine. Can. J. F<strong>or</strong>. Res. 22:1851-1859.<br />

RUDINSKY, J.A. 1966. Host selection and invasion by the Douglas-fir beetle, Dendroctonus pseudotsugae Hopkins, in<br />

coastal Douglas-fir f<strong>or</strong>ests. Can. Entomol. 98: 98-111.<br />

RUMBOLD, C.T. 1936. Three blue-staining fungi, including two new species, associated with bark beetles. J. Agric. Res.<br />

52: 419-437.<br />

SOLHEIM, H. 1988. Pathogenicity of some Ips typographus-associated blue-stain fungi to N<strong>or</strong>way spruce. Medd. N<strong>or</strong>.<br />

inst. skogf<strong>or</strong>sk. 40(14): 1-11.<br />

SOLHEIM, H. 1993. Ecological aspects of fungi associated with the spruce beetle Ips typographus in N<strong>or</strong>way, p. 235-242.<br />

In Wingfield, M.J., Seifert, K.A., and Webber, J.F., eds. Ceratocystis and Ophiostoma Taxonomy, Ecology, and<br />

Pathogenicity. American Phytopathological Society Press, St. Paul, MN.<br />

WHITNEY, H.S. 1982. Relationships between bark beetles and symbiotic <strong>or</strong>ganisms, p. 183-211. In Mitton, J.B. and<br />

Sturgeon, K.B., eds. Bark Beetles in N<strong>or</strong>th American Conifers. University of Texas Press, Austin, TX.<br />

227


INTERRUPTION OF BARK BEETLE AGGREGATION BY A<br />

VIGOR-DEPENDENT PINUS HOST COMPOUND<br />

KENNETH R. HOBSON<br />

Utah State University, Department of F<strong>or</strong>est Resources and <strong>USDA</strong> F<strong>or</strong>est Service, F<strong>or</strong>estry Sciences Lab<strong>or</strong>atroy,<br />

Logan, Utah 84322, USA<br />

INTRODUCTION<br />

The attraction of bark beetles to pheromones has been abundantly demonstrated in the last quarter century (Wood<br />

1982, B<strong>or</strong>den 1985). Much less is known about the response of bark beetles to host volatiles. This paper rep<strong>or</strong>ts preliminary<br />

evidence that one common host compound of N<strong>or</strong>th American conifers is a strong repellent <strong>or</strong> interruptant of aggregation f<strong>or</strong><br />

several species of bark beetles. It also reviews w<strong>or</strong>k that suggests that the level of <strong>this</strong> compound may be an indicat<strong>or</strong> of host<br />

tree stress typically associated with bark beetle infestation in several species of pines. This compound, methyl chavicol,<br />

fulfills the two essential requirements f<strong>or</strong> a kairomonal cue that can explain the selective infestation of susceptible trees by<br />

bark beetles, lit is c<strong>or</strong>related to tree susceptibility, and bark beetles respond to it.<br />

The search f<strong>or</strong> a biochemical indicat<strong>or</strong> of host tree stress that might serve as an olfact<strong>or</strong>y cue f<strong>or</strong> bark beetles is not<br />

new (Rudinsky 1962 and references therein, Cobb et al. 1968). Miller et al. (1968) examined the resin from smog-damaged<br />

ponderosa pines, Pinus ponderosa, in the San Bernardino Mountains east of Los Angeles. These trees were being selectively<br />

infested by mountain pine beetle, Dendroctonus ponderosae, and western pine beetle, Dendroctonus brevicomis. Of the six<br />

monoterpenes examined, none were shown to be significantly different between damaged and healthy trees. However, Cobb<br />

et al. (1972) found a strong, significant difference in the level of one compound, methyl chavicol, in the foliage of smogdamaged<br />

ponderosa pines and healthy trees. Smog-damaged trees had 71% less methyl chavicol in their foliage - by far the<br />

sharpest phytochemical difference found.<br />

M<strong>or</strong>e recently Nebeker et al. (1995) measured 18 host volatiles from healthy lodgepole, Pinus cont<strong>or</strong>ta, and those<br />

diseased by dwarf mistletoe, Arceuthobium americanum; armillaria root disease, Armillaria mellea; <strong>or</strong> comandra blister rust,<br />

Cronartium comandrae. The latter two diseases are rep<strong>or</strong>ted to be the most common predisposing agents of lodgepole to<br />

mountain pine beetle attack in the Intermountain West (Tkacz and Schmitz 1986). In the case of both predisposing diseases<br />

of lodgepole, there was a large and significant difference in the level of 4-allylanisole (a synonym f<strong>or</strong> methyl chavicol)<br />

between healthy and diseased trees. This was not true f<strong>or</strong> dwarf mistletoe, which was not shown to be associated with bark<br />

beetle infestation (Tkacz and Schmitz 1986). F<strong>or</strong> both comandra and armillaria, the difference in the level of methyl chavicol<br />

was one of the greatest phytochemical differences found between diseased and healthy trees. Trees infected with comandra<br />

had 43.6% less methyl chavicol than healthy trees; trees infected with armillaria had 63% less methyl chavicol than healthy<br />

trees. Cobb et al. and Nebeker et al. both found the strongest biochemical differences between diseased and healthy trees in<br />

one compound.<br />

These results begged the question, "What is the response of bark beetles to methyl chavicol?" Electroantennograms<br />

(EAG) show a strong response to methyl chavicol. In 1989, Peter White found that estragole (methyl chavicol) produced the<br />

third highest response of all 11 host compounds tested on the red turpentine beetle, Dendroctonus valens (White and Hobson<br />

1993). In 1993 Armand Whitehead showed a strong EAG response of mountain pine beetle to methyl chavicol (Hobson et al.<br />

in prep). A large EAG response does not indicate that a compound will be a strong attractant <strong>or</strong> repellent. However, ecologically<br />

relevant compounds that produce a large EAG are likely to have behavi<strong>or</strong>al significance (Masson and Mustaparta<br />

1990). These data encouraged us to conduct behavi<strong>or</strong>al field studies.<br />

Mattson, W.J., Niemel_i, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

228


METHODS<br />

In the spring and summer of 1993, we tested the response of five species of bark beetles to methyl chavicol in<br />

Calif<strong>or</strong>nia and Idaho. In April and May, we tested western pine beetle, red turpentine beetle, and two species of Ips, L<br />

paraconfusus and I. pini, at the University of Calif<strong>or</strong>nia's Blodgett F<strong>or</strong>est <strong>Research</strong> <strong>Station</strong> in the central Sierra Nevada.<br />

Tests were conducted as described previously (Hobson 1992, Hobson et al. 1993). F<strong>or</strong> each beetle species, attractive lures<br />

were placed on lindgren traps. In each test there were four treatments" (1) attractant, (2) attractant and methyl chavicol, (3)<br />

methyl chavicol alone, and (4) control blank.<br />

Beetles were collected daily and treatments were randomized after each collection. All four treatments were replicated<br />

in 4-10 blocks. Methyl chavicol was released from four open 1.5-ml epend<strong>or</strong>f tubes at a rate of 0.5 ml/day. The<br />

attractant f<strong>or</strong> D. valens was a set with four 1.5-ml epend<strong>or</strong>f tubes of S(-)[3-pinene. F<strong>or</strong> the western pine beetle, I. pini and I.<br />

paraconfusus commercial lures provided by Phero Tech (Delta, BC) were used. Western pine beetle lures contained exobrevicomin,<br />

frontalin and myrcene. I. pini lures contained ipsdienol (3%+/97%-). I. paraconfusus lures were ipsenol (50%+/<br />

50%-), ipsdienol (97%+/3%-), and cis-verbenol.<br />

In July and August of 1993, we tested mountain pine beetle and I. pini in the Sawtooth National Recreation Area in<br />

south central Idaho. The experimental design was as above. Attractant lures f<strong>or</strong> mountain pine beetle (trans-verbenol, exobrevicomin<br />

and myrcene) were obtained from Phero Tech. Attractant lures f<strong>or</strong> I. pini were as above.<br />

RESULTS<br />

Methyl chavicol significantly reduced attraction of western pine beetle and mountain pine beetle to their respective<br />

aggregation pheromones (Table 1). Attraction of D. valens to [3-pinene was not significantly interrupted by methyl chavicol.<br />

L pini catch was reduced by 29%, but <strong>this</strong> was not significant. I. paraconfusus flew too late in Calif<strong>or</strong>nia f<strong>or</strong> us to collect<br />

adequate data f<strong>or</strong> statistical analysis; however, in five of six blocks where I. paraconfusus was collected, traps baited with<br />

methyl chavicol and the bait had fewer beetles than traps with bait alone.<br />

Table 1.--Bark beetle response to Methyl chavicol<br />

Pheromone Pheromone & Percent<br />

baifi SD methyl chavicol _ SD reduction<br />

Western pine beetle 53.5 a2 69.7 21.6b 28.0 60<br />

Mountain pine beetle 27.4c 47.5 7.9d 12.4 71<br />

Red turpenti ne beetle 10.6 6.28 9.06 6.8 14<br />

Ips pini 20.5 19.7 14.5 12.2 29<br />

_Meancatch.<br />

2Numbersin a row followed by different letters are significantly different o_= 0.05 ANOVA followed by Wilcoxon test.<br />

229


Where beetle response to methyl chavicol was compared by sex, mountain pine beetle males and females were both<br />

significantly less attracted to lures with methyl chavicol. The data f<strong>or</strong> western pine beetle were not sufficient to statistically<br />

test gender-specific response, but methyl chavicol reduced the catch of males by 53% and females by 66%. Methyl chavicol<br />

reduced the catch of D. valens females by 21% and of males by 11%, but neither of these reductions was significant.<br />

DISCUSSION<br />

The aggregation of the two most aggressive bark beetles in <strong>this</strong> study, the mountain and western pine beetles, was<br />

strongly interrupted. The less aggressive species, D. valens, L pini, and I. paraconfusus were not so strongly affected.<br />

Subsequent testing of the L paraconfusus with methyl chavicol and a larger sample size produced a consistent 40%<br />

interruption of aggregation (St<strong>or</strong>er, pers. comm.). The aggregation of two other aggressive bark beetles in Alaska, D.<br />

rufipennis and D. simplex, were interrupted by 82% and 73%, respectively (Werner 1995). Haack and Lawrence have tested<br />

methyl chavicol in traps with attractants of Tomicus piniperda (Haack, pers. comm.). Subsequent to the presentation of <strong>this</strong><br />

paper, Hayes and co-w<strong>or</strong>kers published similar results with Dendroctonusfrontalis (Hayes et al. 1994 a, b) showing 37% and<br />

56% reduction in catch with methyl chavicol (4-allylanisole). Hayes and Strom (1994) also confirmed our results with<br />

mountain pine beetle, obtaining a 77% reduction in catch with methyl chavicol. Their w<strong>or</strong>k differed from <strong>this</strong> w<strong>or</strong>k in that<br />

they could not obtain significant interruption of attraction f<strong>or</strong> western pine beetle, and they did find interruption of attraction<br />

at the 43% level f<strong>or</strong> a Wisconsin population of I. pini.<br />

Methyl Chavicol<br />

Methyl chavicol, an aromatic ether <strong>or</strong> phenylpropanoid, is also commonly known as 4-allylanisole, tarragon, <strong>or</strong><br />

estragol. (Other chemical synonyms include: isoanethole, p-allylmethoxybenzene, 4-allyl-l-methoxybenzene, chavicol<br />

methyl ether, esdragon and 1-methoxy-4-2(2-propenyl)benzene) (Aldrich Chemical - Material Data Safety Sheet). It is<br />

widespread in the ole<strong>or</strong>esin of new w<strong>or</strong>ld pines in the subgenus Pinus, occurring in P ponderosa, P. taeda, P palustris, P.<br />

elliottii, P patula, P jeffreyi, P. tenufolia, P. hartwegii, P michoacana P. lumholtzii (Mirov1961 and references therein), and<br />

P. caribaea (Smith, R.M. 1975). Among old w<strong>or</strong>ld pines it occurs in P. sylvestris and P nigra (Bardysev et al. 1970). It is<br />

abundant in the foliage of ponderosa pine (23%) and present in the foliage of lodgepole and digger pines (Mirov 1961). It is<br />

also found in 12 <strong>or</strong> m<strong>or</strong>e species of spruce (Zavarin, pers. comm.).<br />

Methyl chavicol is the principal volatile ingredient <strong>or</strong> a maj<strong>or</strong> volatile of several strongly scented herbs: tarragon,<br />

Artemisia dracunculus; fennel, Foeniculum vulgare; star anise, Illicium verum; basil, Ocium basilicum; and cloves, Szygium<br />

aromaticum (Duke 1985). Methyl chavicol is also known from the leaves of three rutaceous plants, west african zigua,<br />

Clausena anisata, where it is the maj<strong>or</strong> component of the oil of leaves used to repel mosquitoes (Okunade and Olaifa 1987);<br />

7_nthoxylum spp., where it is toxic to Dacus eggs (Marr and Tang 1992); and southeast asian wood apple, Feronia limonia<br />

(Ahmad et al. 1989). In addition, it is found in oil palm, Elaeis guineensis, where it is attractive to the weevil Elaeidobius<br />

kamerunicus (Hussein et al. 1990). In recent plant herbiv<strong>or</strong>e studies, it is most well known from the w<strong>or</strong>k of Metcalf and his<br />

co-w<strong>or</strong>kers. They have found it in the flowers of Cucurbita maxima, where it is attractive to Diabrotica spp., c<strong>or</strong>n rootw<strong>or</strong>ms.<br />

This association is the basis of a productive research eff<strong>or</strong>t in bi<strong>or</strong>ational management of Diabrotica (Metcalf and<br />

Lampman 1989).<br />

Biosynthetically methyl chavicol is derived from the shikimic acid pathway (Zavarin et al. 1971). This pathway is<br />

disrupted by the herbicide glyphosate (Roundup, Monsanto) (B<strong>or</strong>den, pers. comm.). Interestingly, when Bergvinson and<br />

B<strong>or</strong>den (1991) applied glyphosphate to lodgepole pine, they found that treated trees were readily colonized by mountain pine<br />

beetles. They concluded that the herbicide had inhibited the treated trees' secondary defense response against the beetle's<br />

symbiotic fungi. This is consistent with the hypothesis that the success of mountain pine beetles in glyphosate-treated trees<br />

is, in part, due to a drop in methyl chavicol. Bridges (1987) found methyl chavicol was the most inhibit<strong>or</strong>y host compound to<br />

the three symbiotic fungi associated with D. frontalis, suggesting that it may be an imp<strong>or</strong>tant defensive mechanism of loblolly<br />

pine against southern pine beetle and its vect<strong>or</strong>ed fungi. Hayes et al. (1994a) found that methyl chavicol was strongly<br />

reduced in southern pines following wounding and treatment with metham-sodium and dimethyl sulfoxide. Treated trees<br />

were attacked by D. frontalis after the level of natural methyl chavicol dropped. Methyl chavicol tested subsequently in field<br />

bioassays strongly reduced attraction of southern pine beetle to pheromone-baited traps.<br />

230


The imp<strong>or</strong>tance of a compound as a repellent is somewhat novel in bark beetle host selection research. Limonene<br />

and other terpenes that kill bark beetles and their vect<strong>or</strong>ed fungi have been thought to have a mainly sh<strong>or</strong>t-range effect as<br />

feeding deterrents (Smith, R.H. 1975, Coyne and Lott 1976). The maj<strong>or</strong>ity of w<strong>or</strong>k in <strong>this</strong> field, however, has looked f<strong>or</strong><br />

compounds that increase with stress and are attractive. Few studies have looked f<strong>or</strong> compounds that decrease with stress and<br />

are repellent.<br />

SUMMARY<br />

The significance of these results rests in three principal areas"<br />

1) The large difference in methyl chavicol between healthy and diseased trees provides a clear olfact<strong>or</strong>y sign of stress<br />

f<strong>or</strong> bark beetles, far clearer than has been demonstrated f<strong>or</strong> any monoterpene to date.<br />

2) The stresses that produce the large drop in methyl chavicol are the same stresses that are associated with bark<br />

beetle infestation, suggesting that methyl chavicol is one of the biochemical links between ecology and beetle<br />

behavi<strong>or</strong>.<br />

3) Methyl chavicol reduces the arrestment and landing of bark beetles in the field sufficiently strongly to suggest<br />

management applications f<strong>or</strong> several of our most imp<strong>or</strong>tant bark beetle species with a naturally occurring, abundant<br />

host compound.<br />

Methyl chavicol is the best candidate so far f<strong>or</strong> a compound that responds to stress and to which beetles respond. It<br />

may serve as an indicat<strong>or</strong> and reveal, via biochemical linkage, the elements of host defenses that most strongly affect bark<br />

beetle success. Further investigation of other host species applying stress of the s<strong>or</strong>t known to fav<strong>or</strong> beetle colonization may<br />

determine some new insights f<strong>or</strong> our understanding of the biochemical basis of conifer defenses. Ultimately, we may<br />

understand how environmental stresses that fav<strong>or</strong> beetle colonization produce the characteristics of susceptibility which<br />

beetles fav<strong>or</strong>. We may be able to provide the biochemical linkage between ecological conditions and host selection.<br />

LITERATURE CITED<br />

AHMAD, A., MISRA, L.N., and THAKUR, R.S. 1989. Composition of the volatile oil from Feronia limonia leaves.<br />

Planta-Medica 55" 199-200.<br />

BARDYSEV, I.I., PAPANOV, G.J., and PERCOVSKIJ, A.L. 1970. Chemical composition of the balsams ofPinus sylvestris<br />

and P. nigra growing in Bulgaria. Dokl. AN BSSR 14: 539-40.<br />

BERGVINSON, D.J. and BORDEN, J.H. 1991. Glyphosate-induced changes in the attack success and development of the<br />

mountain pine beetle and impact of its natural enemies. Entomologia Experimentalis et Applicata. 60:203-212.<br />

BORDEN, J.H. 1985. Aggregation pheromones. In Kerkut, G.A. and Gilbert, L.I., eds. Comprehensive Insect Physiology,<br />

Biochemistry & Pharmacology. Pergamon Press, Oxf<strong>or</strong>d.<br />

BRIDGES, J.B. 1987. Effects of terpenoid compounds on growth of symbiotic fungi associated with the southern pine<br />

beetle. Phytopathol. 77: 83-85.<br />

COBB, EW. Jr., WOOD, D.L., STARK, R.W., and MILLER, P.R. 1968. The<strong>or</strong>y on the relationships between oxidant injury<br />

and bark beetle infestation. Hilgardia 39: 141-152.<br />

COBB, EW. Jr., ZAVARIN, E., and BERGOT, J. 1972. Effect of air pollution on the volatile oil from leaves of Pinus<br />

ponderosa. Phytochemistry 11: 1815-1818.<br />

COYNE, J.F. and LOTT, L.H. 1976. Toxicity of substances in Pine ole<strong>or</strong>esin to southern pine beetles. J. Ge<strong>or</strong>gia Entomol.<br />

Soc. 11: 301-305.<br />

231


DUKE, J.A. 1985. CRC Handbook of Medicinal Herbs. CRC Press, Boca Raton, FL. 677 p.<br />

HAYES, J.L., INGRAM, L.L. Jr., STROM, B.L., ROTON, L.M., BOYETTE, M.W., and WALSH, M.T. 1994a. Identification<br />

of a host compound and its practical application: 4-allylanisole as a bark beetle repellent. Proceedings of the 4th<br />

Southern <strong>Station</strong> Chemical Sciences Meeting; Feb. 1-2, 1994. Gen. Tech. Rep. SO-104. Asheville, NC: U.S.<br />

Department of Agriculture, F<strong>or</strong>est Service.<br />

HAYES, J.L., STROM, B.L., ROTON, L., and INGRAM, L.L. Jr. 1994b. Repellent properties of the host compound 4allylanisole<br />

to the southern pine beetle. J .Chem. Ecol. p. 1595-1615.<br />

HAYES, J.L. and STROM, B.L. 1994. 4-Allylanisole as an inhibit<strong>or</strong> of bark beetle (Coleoptera: Scolytidae) aggregation. J.<br />

Econ. Entomol. 87: 1586-1594.<br />

HOBSON, K.R. 1992. Studies on host selection and biology of Dendroctonus valens and fungal host interactions of<br />

Dendroctonus brevicomis. Ph.D. Dissertation, University of Calif<strong>or</strong>nia, Berkeley.<br />

HOBSON, K.R,, WOOD, D.L., COOL, L.G., WHITE, RR., OHTSUKA, T., KUBO, I., and ZAVARIN, E. 1993. Chiral<br />

specificity :inresponses by the bark beetle D. valens to host kairomones. J. Chem. Ecol. 19: 1837-1846.<br />

HUSSEIN, M.Y., LABS, N.H., and ALI, J.H. 1990. Biological and chemical fact<strong>or</strong>s associated with the successful introduction<br />

of Elaeidobius kamerunicus Faust, the oil palm pollinat<strong>or</strong> in Malaysia. In van Heemert, C. and de Ruijter, A.,<br />

eds., The Sixth Int'l Syrup. on Pollination.<br />

MARR, K.L. and TANG, C.S. 1992. Volatile insecticidal compounds and chemical variability of Hawaiian Zanthoxylum<br />

(Rutaceae) species. Biochem. System. and Ecol. 20: 20%217.<br />

MASSON, C. and MUSTAPARTA, H. 1990. Chemical inf<strong>or</strong>mation processing in the olfact<strong>or</strong>y system of insects. Physiol.<br />

Rev. 70: 199-245.<br />

METCALF, R.L. and LAMPMAN, R.L. 1989. Estragole analogues as attractants f<strong>or</strong> Diabrotica species (Coleoptera:<br />

Chrysomelidae) c<strong>or</strong>n rootw<strong>or</strong>ms. J. Econ. Entomol. 82: 123-129.<br />

MILLER, RR., COBB, EW., Jr., and ZAVARIN, E. 1968. Effect of injury upon ole<strong>or</strong>esin composition, phloem carbohydrates,<br />

and phloem pH. Hilgardia 39: 135-139.<br />

MIRO'v; N.T. 1961. Composition of gum turpentines of pines. Tech. Bull. 1239. Washington, De: U.S. Department of<br />

Agriculture.<br />

NEBEKER, T.E., SCHMITZ, R.E, TISDALE, R.A., and HOBSON, K.R. 1995. Chemical and nutritional status of dwarf<br />

mistletoe, armillaria root rot, and comandra blister rust infected trees that may influence tree susceptibility to bark<br />

beetle attack. Can. J. Bot. (accepted 11/94, in press)<br />

OKUNADE, A.L. and OLAIFA, J.I. 1987. Estragole: an acute toxic principle from the volatile oil of the leaves of Clausena<br />

anisata. J. Nat. Prod. 50: 990-991.<br />

RUDINSKY, J.A. 1962. Ecology of Scolytidae. Ann. Rev. Entomol. 7: 327-348.<br />

SMITH, R.H. 1975. F<strong>or</strong>mula f<strong>or</strong> describing effect of insect and host tree fact<strong>or</strong>s on resistance to western pine beetle attack.<br />

J. Econ. Entomol. 68(6): 841-844.<br />

SMITH, R.M. 1975. Note on gum turpentine of Pinus caribaea from Fiji. N. Zealand J. of Sci. 18: 547-548.<br />

TKACZ, B.M. and SCHMITZ, R.E 1986. Association of an endemic mountain pine beetle population with lodgepole pine<br />

infected by ArmiUaria root disease in Utah. Res. Note INT-353. Ogden, LIT: U.S. Department of Agriculture, F<strong>or</strong>est<br />

Service.<br />

232


WERNER, R.A. 1995. Toxicity and repellency of 4-allylanisole and monoterpenes from white spruce and tamarack to the<br />

spruce beetle and eastern larch beetle (Coleoptera: Scolytidae) Environ. Entomol. 24: 151-158.<br />

WHITE, RR. and HOBSON, K.R. t993. Stereospecific antennal response by the red turpentine beetle, Dendroctonus valens<br />

to chiral monoterpenes from ponderosa pine resin. J. Chem. Ecol. t9:2193-2202.<br />

WOOD, D.L. 1982. The role of pheromones and atlomones in the host selection and colonization behavi<strong>or</strong> of bark beetles.<br />

Ann. Rev. Entomol. 27:411-446.<br />

ZAVARIN, E., COBB, F.W., Jr., BERGOT, J. and BARBER, H.W. 1971. Variation of the Pinus ponderosa needle oil with<br />

season and needle age. Phytochemistry 10:3107-3114.<br />

233


WILL GLOBAL WARMING ALTER PAPER BIRCH<br />

SUSCEPTIBILITY TO BRONZE BIRCH BORER ATTACK?<br />

ROBERT A. HAACK<br />

<strong>USDA</strong> F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>est Experiment <strong>Station</strong><br />

1407 S. Harrison Road, East Lansing, M148823, USA<br />

INTRODUCTION<br />

Acc<strong>or</strong>ding to several general circulation models, a rapid climatic warming of 1-5°C in the annual global-mean surface<br />

air temperature is predicted by the year 2050 (Shands and Hoffman 1987, MacCracken et al. 1990, Karl et al. 1991,<br />

Schlesinger and Jiang 1991, IPCC 1992, Peters 1992, Bengtsson 1994). Some f<strong>or</strong>est entomologists predict an increase in<br />

outbreak frequency of certain f<strong>or</strong>est insects as a result of global warming (Kellomaki et al. 1988, Hedden 1989, Franklin et<br />

al. 1992, Fleming and Volney 1995), although the actual response may be difficult to predict due to the many interacting<br />

biotic and abiotic fact<strong>or</strong>s (Hedden 1989, P<strong>or</strong>ter et al. 1991, Cammell and Knight 1992, Dewar and Watt 1992, Ayres 1993,<br />

Scriber and Gage 1995, Williams and Liebhold 1995).<br />

Is there a way to test whether climate change will alter tree susceptibility to insect attack? One method is to look f<strong>or</strong><br />

patterns of differential insect attack within genetic test plantations where trees of known genetic background grow together at<br />

the same test site. Frequently, the primary objective of genetic test plantings is to identify either specific families <strong>or</strong> geographic<br />

seed sources (provenances) that perf<strong>or</strong>med well at a particular location, especially with respect to growth and f<strong>or</strong>m<br />

characteristics. The trees in genetic test plantings, particularly provenance tests, commonly <strong>or</strong>iginate from several distant<br />

locations, often representing several contrasting climates. Because tree populations become genetically adapted to their local<br />

conditions, one maj<strong>or</strong> fact<strong>or</strong> that affects their perf<strong>or</strong>mance when moved to a new location is the difference in climate between<br />

the <strong>or</strong>iginal and the new location (Wright 1976, Kozlowski et al. 1991, Schmidtling 1994). Data from genetic test plantings<br />

have only recently been examined to elucidate how some tree species might respond to global warming (Matyas 1994,<br />

Schmidtling 1994). Similarly unexpl<strong>or</strong>ed has been the use of provenance plantation data to predict how global warming will<br />

affect tree susceptibility to insect attack.<br />

In <strong>this</strong> paper, I expl<strong>or</strong>e how global warming might alter susceptibility of paper birch, Betula papyrifera Marsh., to the<br />

bronze birch b<strong>or</strong>er, Agrilus anxius G<strong>or</strong>y (Coleoptera: Buprestidae). This particular tree-insect combination was selected<br />

because (1) paper birch appears highly sensitive to environmental stress given that several large scale declines of paper birch<br />

have been rep<strong>or</strong>ted <strong>this</strong> century in eastern N<strong>or</strong>th America, and (2) the bronze birch b<strong>or</strong>er has usually been the ultimate<br />

m<strong>or</strong>tality agent of stressed birch trees (Slingerland 1906, Swaine 1918, Spaulding and MacAloney 1931, Balch and Prebble<br />

1940, Hawboldt 1947, Nash et al. 1951, Barter 1957, Redmond 1957, Clark and Barter 1958, Haack and Mattson 1989, Jones<br />

et al. 1993, Braathe 1995). Birch m<strong>or</strong>tality as a result of bronze birch b<strong>or</strong>er attack has followed stress events such as drought,<br />

extreme cold winter temperatures that follow closely behind a winter thaw, elevated growing season temperatures, high soil<br />

water tables, years of heavy seed crops, and years of severe insect defoliation (Redmond 1955, 1957; Clark and Barter 1958;<br />

Herms 1991; Herms and Mattson 1991; Auclair et al. 1992; Jones et al. 1993; Auclair et al. 1995; Braathe 1995).<br />

Paper birch is a b<strong>or</strong>eal species, and in Michigan it reaches its southernmost range near the center of Michigan's lower<br />

peninsula. The ecotone between b<strong>or</strong>eal and temperate f<strong>or</strong>ests closely follows the 47°F (8.3"C) average annual mean temperature<br />

isotherm in Michigan (Fig. 1). Typically, annual mean temperature decreases steadily with increasing latitude, but in<br />

Michigan, the Great Lakes cause a semi-marine type climate, causing isotherms to closely follow the lakes' sh<strong>or</strong>elines (Fig.<br />

1). This feature of moderately vertical isotherms in p<strong>or</strong>tions of Michigan allows f<strong>or</strong> great variation in climatic conditions<br />

within a rather narrow latitudinal band. Michigan is relatively flat, ranging in elevation from 174 to 604 m (572-1,980 ft)<br />

above sea level.<br />

Mattson, W.J., Niemel/i, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

234


41 40<br />

42 \ 4t<br />

42 41<br />

42 ",,,\<br />

42 40 40<br />

41 / 42<br />

40 43<br />

4O<br />

41,1 44<br />

2<br />

43<br />

44<br />

42 45 ,,.<br />

45<br />

46 /<br />

47 ,_ 46<br />

48 I 1<br />

49 _ 49<br />

.,, /<br />

49 49 50<br />

Figure 1.--Average annual mean temperature isotherms f<strong>or</strong> Michigan in °E based on the period 1940-1969 (MDA 1974).<br />

Inset shows Michigan, with county-level resolution, the surrounding Great Lakes (shaded area), and outlines of<br />

several adjacent states and Canada. Approximate °C equivalent values are: 40°F = 4.4°C, 41 °F = 5.0°C, 42°F = 5.60C,<br />

43°F = 6.1 °C, 44°F = 6.7°C, 45*F = 7.2°C, 46°F = 7.7°C, and 47°F = 8.3°C<br />

Models of global warming in the n<strong>or</strong>thern hemisphere predict that a species range will change most dramatically<br />

along the southern limits of its range (Cannell et al. 1989, Botkin and Nisbet 1992, Woodward 1992, Matyas 1994, Sykes and<br />

Prentice 1995). F<strong>or</strong>tunately, in southern Michigan there is a paper birch progeny test (Fig. 2) that consists of over 200 halfsib<br />

(i.e., where the identity of the maternal parent is known) families that were collected from throughout the natural range of<br />

birch in Michigan. In addition, it is coincidental that the mean annual temperature varies by about 10°F (5.6°C) between the<br />

southern and n<strong>or</strong>thern extremes of Michigan (Fig. 1). Theref<strong>or</strong>e the paper birch trees growing in the genetic test plantation in<br />

southern Michigan are the<strong>or</strong>etically experiencing an increase in mean annual temperature of 1° to 5°C. Given that the bronze<br />

birch b<strong>or</strong>er preferentially attacks weakened birch trees, it seems plausible that b<strong>or</strong>er attack could vary among seedlots of<br />

varying geographic <strong>or</strong>igins if they were experiencing differential levels of stress.<br />

The bronze birch b<strong>or</strong>er, like many other buprestids in the genus Agrilus, usually attacks host trees that are stressed by<br />

drought <strong>or</strong> repeated insect defoliation (Anderson 1944; Clark and Barter 1958; Barter 1965; Carlson and Knight 1969; Haack<br />

and Benjamin 1982; Haack and Slansky 1987; Mattson and Haack 1987a, 1987b; Haack and Mattson 1989; Herms 1991;<br />

Herms and Mattson 1991; Wargo and Haack 1991; Haack 1992; Haack and Acciavatti 1992; Solomon 1995). The biology<br />

and life hist<strong>or</strong>y of the bronze birch b<strong>or</strong>er has been rep<strong>or</strong>ted by several auth<strong>or</strong>s (Chittenden 1898; Balch and Prebble 1940;<br />

Anderson 1944; Barter 1957; Ball and Simmons 1980; Loerch and Cameron 1983, 1984; Akers and Nielsen 1990; Wilson<br />

and Haack 1990). Briefly, the bronze birch b<strong>or</strong>er is univoltine. Adult beetles emerge in early summer from within their host 235<br />

50<br />

..,- 47


47°N - /<br />

8 1<br />

45ON _<br />

44ON _<br />

43ON_<br />

42°N<br />

Figure 2.--State of Michigan, showing the 26 locations where paper birch seed was collected in 1975. The * in southwestern<br />

Michigan indicates the location of the plantation site where the seedlings were planted in 1976. Inset shows the<br />

conterminous 48 states of the US, with Michigan shaded.<br />

trees, and then search f<strong>or</strong> suitable trees to attack. Adults often reattack the same tree until it dies, but from year to year they<br />

attack only those p<strong>or</strong>tions that are still living at the time of oviposition. It often takes 2-3 years of successive attack bef<strong>or</strong>e a<br />

given branch <strong>or</strong> trunk section dies. Adults are believed to be attracted primarily by host od<strong>or</strong>s given that no pheromone has<br />

yet been discovered f<strong>or</strong> a buprestid (Haack and Slansky 1987, Haack 1992). After mating, females oviposit on the bark<br />

surface of the trunk and branches. After hatching, larvae tunnel through the bark and then feed in the cambial region, sc<strong>or</strong>ing<br />

both the inner bark and outer sapwood. Larvae construct meandering galleries throughout the summer months, passing<br />

through four larval instars. In late summer <strong>or</strong> fall, larvae construct individual pupal chambers in the outer sapwood and there<br />

pass the winter. Pupation occurs the following spring, and soon the adults emerge to renew the cycle.<br />

MATERIALS AND METHODS<br />

As described in Miller et al. (1991), seed was collected from 218 open-pollinated paper birch seedlots (half-sib<br />

families) in 1975, representing 26 Michigan populations; 123 trees from 13 populations in Michigan's Upper Peninsula and<br />

95 trees from 13 populations in Michigan's Lower Peninsula (Fig. 2, Table 1). All parents were growing in natural f<strong>or</strong>ests<br />

between latitudes 43.3°N and 47.4°N, and ranged from 5 to 22 m in height, and 9 to 46 cm in diameter (Table 1). The seed<br />

from the 218 half-sib families was planted in February 1976 and grown in the greenhouse f<strong>or</strong> 5 months, held outdo<strong>or</strong>s f<strong>or</strong> 2<br />

236


Table 1.--Summary data f<strong>or</strong> the 26 populations of Michigan paper birch trees, representing 218 trees, from which seed was<br />

collected in 1975, including f<strong>or</strong> each population the number of trees sampled, location data, climatic data, and mean<br />

height and diameter at breast height (DBH) of the parental trees.<br />

Seed source I Location 2 Climatic data3 Height (m) 4 DBH (cm) 4<br />

Pop No °N Lat Area County Temp DD FFD Mean (range) Mean (range)<br />

t 7 46.8 UP Ontonagon 42 1600 I10 14.0 (10-21) 25.3 (9-46)<br />

2 10 47.4 UP Keweenaw 40 1600 140 11.8 (5-19) 24.0 (10-45)<br />

3 10 46.7 UP Baraga 40 1600 130 16.2 (9-21) 21.0 (15-29)<br />

4 7 46.3 UP Gogebic 40 1800 60 15.0 (12-19) 22.6 (21-24)<br />

5 10 46.1 UP Iron 41 1800 100 15.1 (13-16) 19.6 (15-25)<br />

6 10 45.2 UP Menominee 42 1600 140 14.6 (12-16) 18.6 (14-23)<br />

7 10 46.1 UP Delta 41 1800 70 13.4 (9-16) 20.4 (13-31)<br />

8 10 46.6 UP Marquette 42 1600 150 14.7 (8-21) 23.5 (10-40)<br />

9 10 46.5 UP Alger 41 1400 140 12.0 (11-15) 16.5 (13-20)<br />

10 11 46_0 UP Schoolcraft 42 1600 110 16.3 (12-22) 24.6 (14-32)<br />

11 9 46.7 UP Chippewa 41 1400 150 11.1 (11-13) 17.6 (12-28)<br />

12 10 46.3 UP Chippewa 41 1400 130 13.5 (11-16) 22.6 (17-30)<br />

13 9 46.0 UP Mackinac 41 1600 120 13.2 (8-16) 18.3 (15-24)<br />

14 10 45.6 LP Emmet 42 1800 150 13.7 (7-16) 25.4 (15-32)<br />

15 5 44.9 LP Antrim 44 2000 140 15.5 (14-16) 24.5 (21-26)<br />

16 10 44.9 LP Montm<strong>or</strong>ency 43 1800 70 13.7 (11-16) 19.8 (13-25)<br />

17 10 45.5 LP Presque Isle 44 1800 140 12.3 (9-16) 17.7 (12-28)<br />

18 10 44.8 LP Alcona 43 1800 130 16.3 (15-18) 25.8 (19-33)<br />

19 4 43.4 LP Montcalm 46 2400 130 15.2 (11-18) 21.5 (14-27)<br />

20 5 43.9 LP Lake, Mason 46 2200 120 14.6 (11-16) 21.7 (19-24)<br />

21 8 44.3 LP Manistee 46 2200 150 15.8 (11-19) 20.5 (16-27)<br />

22 9 44.4 LP Missaukee 43 1800 100 14.5 (11-19) 22.6 (14-38)<br />

23 9 44.2 LP Gladwin 44 2000 100 16.2 (11-18) 20.4 (15-28)<br />

24 9 44.0 LP Arenac 45 2000 130 13.8 (11-16) 17.5 (11-25)<br />

25 5 43.6 LP Sanilac 46 2400 130 12.4 (11-15) 15.6 (13-17)<br />

26 1 43.3 LP Saginaw 47 2600 150 12.1 n.a. 18.2 n.a.<br />

Seed source: Pop - seed source population number as given in Fig. 1; No = number of parental trees from which seed<br />

was collected in each particular population.<br />

2 Location: Area - Michigan seed source locations where UP refers to Upper Peninsula and LP refers to Lower<br />

Peninsula; County = Michigan county from which trees were sampled.<br />

3Climatic data f<strong>or</strong> each population of sampled trees using isopleth data from various Michigan climate maps: Temp =<br />

average annual mean temperature (°F; 1940-1969; MDA 1974); DD - seasonal (March - October) growing degree days<br />

(base 50°F; 1951-1980; Eichenlaub et al. 1990); FFD - average number of frost free days (i.e., number of days between<br />

last spring 32°F reading and first fall 320F reading; 1940-1969; data drawn from MDA 1971).<br />

4 Height and diameter at breast height (DBH; 1.3 m) of sampled trees by population.<br />

weeks, and then planted in a plowed and disked field in June, 1976. The plantation is located near 42°N, at the Fred Russ<br />

Experimental F<strong>or</strong>est, Michigan State University, in southwestern Michigan (Fig. 2). Trees were planted in a completely<br />

randomized design on 2.4 x 2.4 m square spacing with an average of 5 trees per seedlot. Overall, 1,088 trees were planted,<br />

with 636 trees from Upper Peninsula seed sources and 452 trees from Lower Peninsula seed sources.<br />

Measurements of <strong>this</strong> plant material began in 1976 and continued through 1992. Height, diameter, bark col<strong>or</strong>, insect<br />

damage, and tree survival were measured at varying intervals over 15 years (see also Miller et al. 1991). Height was measured<br />

in years 2, 6, 7, 8, and 10 post-establishment. Bronze birch b<strong>or</strong>er damage was evaluated in years 10, 12, and 15 on a<br />

scale of 0 to 3, where 0 = tree apparently healthy with no evidence of b<strong>or</strong>er infestation; 1 = less than half of the crown<br />

237


anches dead <strong>or</strong> dying with b<strong>or</strong>er-related bark ridges obvious on the bark surface; 2 = m<strong>or</strong>e than half of the crown branches<br />

dead <strong>or</strong> dying with bark ridges obvious and b<strong>or</strong>er exit holes often present; and 3 = tree dead, usually with b<strong>or</strong>er-related bark<br />

ridges and exit holes present. Bark ridges (callus tissue) occur on branch and trunk sections in response to larval feeding<br />

(Anderson 1944, Barter 1957, Wilson and Haack 1990, Herms 1991). The presence of b<strong>or</strong>er-related bark ridges was sc<strong>or</strong>ed<br />

in years 12 and 15. Completion of larval development was also sc<strong>or</strong>ed in years 12 and 15 by noting the presence of b<strong>or</strong>er exil<br />

holes. Tree survival was computed f<strong>or</strong> each of the 218 families in years 2, 6, 8, 10, 12, 15, and 16.<br />

Hist<strong>or</strong>ical temperature data were obtained from the Michigan Department of Agriculture, Climatological Division<br />

(MDA 1974). Average annual mean temperature was considered the most relevant climatic variable to use when considering<br />

global warming. The average annual mean temperature assigned to each seed source was obtained from a map showing<br />

annual mean temperatures f<strong>or</strong> Michigan averaged over the 30-year period 1940-1969 (Fig. 1, Table 1, MDA 1974). This was<br />

done by locating all seed collection sites on the map and then assigning a particular annual mean temperature value to each<br />

seed source based on its location relative to the map isotherms. Although actual temperature data f<strong>or</strong> each of the 26 collections<br />

sites would have been preferred, <strong>this</strong> was not possible because many collection sites had no nearby weather stations.<br />

Additional climatic data f<strong>or</strong> each of the 26 collection sites were obtained in similar fashion from other maps (Table 1).<br />

Using the map shown in Figure 1, the annual mean temperature values assigned to the 218 seed sources ranged from<br />

40° to 47°F (4.4*-8.3°C; Table 1). Actual average annual mean temperature values from several official weather stations f<strong>or</strong><br />

<strong>this</strong> same area and time period ranged from 39.7* to 47.1"F (4.30-8.4"C; MDA 1971). The average annual mean temperature<br />

of the plantation site was about 49.00F (9.4°C; MDA 1971). Overall, assuming that the birch seed sources were genetically<br />

acclimated to the average annual mean temperature of their <strong>or</strong>iginal sites, then the birch trees growing at the plantation site<br />

were the<strong>or</strong>etically experiencing levels of climatic warming equivalent to 1.9° to 9.3*F (1.1 ° to 5.2°C; Table 2). F<strong>or</strong> ease of<br />

discussion, I will use the term climatic differential (CD) to represent the difference (in °C) between the average annual mean<br />

temperatures of the seedlot <strong>or</strong>igin and the plantation site.<br />

Table 2.--Average annual mean temperature data and percent of paper birch trees (both live and dead) that had bronze<br />

birch b<strong>or</strong>er-induced ridges and/<strong>or</strong> b<strong>or</strong>er exit holes along their lower trunk in years 12 (1987) and 15 (1990) postestablishment<br />

by climatic differential treatment.<br />

Mean temperature<br />

difference between<br />

C<strong>or</strong>responding <strong>or</strong>igin and Percent of trees with Percent of trees with<br />

isotherms 2 plantation site 3 b<strong>or</strong>er-induced ridges 4 b<strong>or</strong>er exit holes 4<br />

CD j °F °C *F °C Yr 12 Yr 15 Yr 12 Yr 15<br />

CD t ° 46, 47 7.7, 8.3 2 1.1 30ab 32b 10b 13b<br />

CD2 ° 44, 45 6.7, 7.2 4 2.2 24b 33b 8b 18b<br />

CD3 ° 42, 43 5.6, 6.1 6 3.3 39ab 46b 12b 25b<br />

CD4 ° 40, 41 4.4, 5.0 8 4.4 50a 60a 23a 40a<br />

CD = Climatic differential, which is a term that reflects the difference between the average annual mean temperature<br />

of the seed sources within a particular treatment group and the average annual mean temperature of the plantation<br />

site, which was 49°F (9.4"C). The value 1° in CD1 °, f<strong>or</strong> example, reflects the annual mean temperature difference<br />

between the <strong>or</strong>igin and the plantation site, e.g., 1.I*C (see footnote 3 below).<br />

2The isotherms of average annual mean temperature in *F (and the approximate °C equivalent), as given in Fig. 1,<br />

that c<strong>or</strong>respond to each CD grouping.<br />

3F<strong>or</strong> each CD grouping, the mean temperature difference between the <strong>or</strong>igin and the plantation site was calculated<br />

as follows. F<strong>or</strong> CDI*, the c<strong>or</strong>responding isotherms were 46* and 47°F. However, these isotherms span the temperature<br />

range from 46* to 47.9°F, with a mean of about 47"E Thus, the temperature difference between the <strong>or</strong>igin and the :<br />

plantation site in <strong>this</strong> case is 2*F (49°-47 *) <strong>or</strong> about 1.I*C. These values represent the the<strong>or</strong>etical level of climatic<br />

warming that the c<strong>or</strong>responding trees were experiencing while growing at the plantation site in southern Michigan.<br />

4 Percent of trees with b<strong>or</strong>er-induced ridges and/<strong>or</strong> b<strong>or</strong>er exit holes along the lower trunk; percent values followed by<br />

the same letter within a column are not significantly different<br />

followed by the Dunn's test f<strong>or</strong> multiple comparisons).<br />

238<br />

(p


One weakness of using data from genetic tests plantings to study the impact of climatic warming, however, is that<br />

other variables besides temperature change when moving seed from its site of <strong>or</strong>igin to the test location. Photoperiod, f<strong>or</strong><br />

example, is one imp<strong>or</strong>tant variable f<strong>or</strong> which many tree species, including paper birch, are genetically adapted (Vaartaja<br />

1959, Wright 1976, Kozlowski et al. 1991, Farmer 1993). In the n<strong>or</strong>thern hemisphere, typically, n<strong>or</strong>thern sources tend to<br />

flush earlier and terminate shoot growth earlier when moved southward compared with m<strong>or</strong>e local sources (Wright 1976,<br />

Kozlowski et al. 1991). In the present study, birch seed was collected between Lat. 43.3°N and 47.4°N, and then planted at<br />

42.0°N. On the longest day of the year there is about 15.2 hours of light at 42.0°N, 15.5 hours at 43.3°N, and 16.0 hours at<br />

47.4°N. Similarly, on the sh<strong>or</strong>test day of the year, they experience about 9.1 hours of light at 42.0°N, 9.0 hours at 43.3°N, and<br />

8.4 hours at 47.4°N. As a means of minimizing any influence of photoperiod, linear regression analyses were also conducted<br />

on two subsets of the <strong>or</strong>iginal 26 populations (collection sites) in which the latitudinal spread among the seed sources was<br />

further reduced. In the first case, only the 13populations from the Lower Peninsula of Michigan were tested (populations 14-<br />

26, Fig. 2, Table 1), representing collections within a 2° span of latitude and a 5°F (2.8°C) span in annual mean temperature.<br />

In the second case, 12 birch populations were tested (populations 4-7, 10, 12-18; see Fig. 2, Table 1), representing collections<br />

within a 1.5° span in latitude and a 4°F (2.2°C) span in annual mean temperature.<br />

Analyses were conducted using SAS and SigmaStat. Height and b<strong>or</strong>er data were analyzed on an individual tree basis,<br />

whereas tree survival was calculated on a family basis (n = 218), using only those birch families that had at least one tree<br />

surviving through the second growing season (1977). Each tree was assigned to one of four "treatments" based on the<br />

average annual mean temperature of its <strong>or</strong>igin. The four groupings were isotherms 40 ° and 41°F, 42 ° and 43°F, 44 ° and 45°F,<br />

and 46 oand 47°F (Table 2). The climatic differential term assigned to each grouping was CD1 ° (46-47°F), CD2°(44°-45°F),<br />

CD3 ° (42°-43°F), and CD4 ° (40°-41°F; Table 2). These terms, CDI°-CD4 °, reflect the average climatic differential in °C fbr<br />

each treatment group, e.g., f<strong>or</strong> CD1 ° the midpoint climatic differential value is actually 1.I°C and represents the average<br />

degree of climatic warming that the trees in isotherms 460F and 47°F were the<strong>or</strong>etically experiencing (Table 2). Analysis of<br />

variance (ANOVA) was used to test f<strong>or</strong> differences among treatments; ap level of 0.05 was used f<strong>or</strong> significance. Tree<br />

height and survival data were analyzed using the GLM procedure of SAS, followed by the Student-Newman-Keuls means<br />

separation test. The arcsin square root transf<strong>or</strong>mation was perf<strong>or</strong>med on the survival data pri<strong>or</strong> to analysis. B<strong>or</strong>er data were<br />

analyzed using the Kruskal-Wallis nonparametric test of SigmaStat, which perf<strong>or</strong>ms an ANOVA on ranks; Dunn's test was<br />

used to make multiple comparisons among treatment means.<br />

RESULTS<br />

Paper birch growth, survival, and susceptibility to bronze birch b<strong>or</strong>er attack were significantly affected by the<br />

climatic differential between the <strong>or</strong>igin of the seed source and the plantation site (Fig. 3). Overall, height growth decreased<br />

as the climatic differential increased (Fig. 3A). Differences in height growth were already evident by the end of the second<br />

growing season in 1977 (p < 0.0001; Fig. 3A). By the end of the sixth growing season (1985), CDI ° trees were the tallest,<br />

with tree height becoming successively sh<strong>or</strong>ter f<strong>or</strong> the CD2 °, CD3 °, and CD4 ° trees (p < 0.0001 ; Fig. 3A).<br />

As the trees aged, tree survival tended to decrease as the climatic differential increased (Fig. 3B). From year 2 (1977)<br />

to year 8 (1983) post-establishment, little tree m<strong>or</strong>tality occurred, but thereafter tree m<strong>or</strong>tality accelerated, especially among<br />

the trees experiencing the greatest climatic differential (CD4°; Fig. 3B). Treatment survival values first became significantly<br />

different in year 12 post-establishment (1987; p < 0.006), when survival f<strong>or</strong> the CD4 ° trees was lower than that f<strong>or</strong> trees from<br />

the other three groups. A similar pattern of tree survival continued into years 15 (1990; p < 0.001) and 16 (1991; p < 0.0008)<br />

post-establishment.<br />

As the trees became older, tree susceptibility to the bronze birch b<strong>or</strong>er increased with increasing climatic differential<br />

(Fig. 3C). Of the 146 paper birch trees that died between 1985 and 1991, 143 (98%) of them had signs of b<strong>or</strong>er attack along<br />

their lower trunk. A similar pattern emerged during each of the 3 years when b<strong>or</strong>er damage was assessed (1985, 1987, 1990;<br />

p < 0.001 f<strong>or</strong> each), i.e., CD4 ° trees experienced the highest levels of b<strong>or</strong>er attack, while CDI ° trees experienced the lowest<br />

levels (Fig. 3C). As would be expected from the above discussion, the occurrence of b<strong>or</strong>er-related bark ridges and exit holes<br />

was m<strong>or</strong>e common on CD4 ° trees than on CD 1° trees (Table 2).<br />

Regarding the two subsets of birch populations in which the latitudinal spread was reduced, paper birch growth,<br />

survival, and susceptibility to bronze birch b<strong>or</strong>er were all significantly affected by the climatic differential between the <strong>or</strong>igin<br />

of the seed source and the plantation site. In the first case, linear regression analysis of the 13 populations from Michigan's<br />

239


- A<br />

7 A<br />

_"<br />

--l--<br />

6<br />

5<br />

C<br />

c_ -,.-- CD1 o<br />

"1-<br />

U_I<br />

3 _ CD2 °<br />

I'-- r'r"<br />

i"Ti<br />

2<br />

4 jc<br />

_ CD3 o<br />

1 I_B +- CU4 °<br />

__1<br />

! r 1 _ _ _ .... f i' _ ...... _ T ' i ' E 1<br />

2 4 6 8 10 12 14 16<br />

100 : ........... A A<br />

0_ _ 70 ",_ ",,,_AA..<br />

,,u4,6o<br />

¢7" B B<br />

!--'- 50<br />

40 , ," , ..........<br />

2 4 6 8 10 12 14 16<br />

"_0_"/ , .... .., ...., ..,B , ...., , ., , ,<br />

2 4 6 8 10 12 14 16<br />

TREE AGE (Years)<br />

Figure 3.kPaper birch mean height growth (A), mean survival (B), and mean susceptibility to b<strong>or</strong>er attack (C) over time<br />

(years post-establishment) in a provenance plantation in southern Michigan. Survival was based on only those trees<br />

that lived f<strong>or</strong> the first two years. The 0-3 b<strong>or</strong>er rating scale is presented in the Methods section. Trees are divided<br />

into four treatment groups based on the climatic differential (CD) between the average annual mean temperature of<br />

the <strong>or</strong>igin of each seed source and the annual mean temperature of the plantation site. On average, trees assigned to<br />

treatment CD 1° had a 1.1°C difference between the <strong>or</strong>igin and the test site, CD2 ° a 2.2°C difference, CD3 ° a 3.3°C<br />

difference, and CD4 ° a 4.4*C difference (see Table 2). F<strong>or</strong> each parameter, means followed by the same letter within<br />

a column (i.e., year) are not significantly different (p


Lower Peninsula (populations 14-26) demonstrated that as the climatic differential between the <strong>or</strong>igin and the test site<br />

increased, tree height decreased (p < 0.006, r2 = 0.51, n = 13 birch populations), tree survival decreased (p < 0.012, r2 = 0.49,<br />

n = 13), and tree susceptibility to b<strong>or</strong>er attack increased (p < 0.006, r2= 0.51, n = 13). Similarly, in the second case, which<br />

included 12 birch populations (populations 4-7, 10, 12-18), linear regression analysis indicated that as the climatic differential<br />

between <strong>or</strong>igin and test site increased, tree height decreased (p < 0.02, r2= 0.43, n = 12 birch populations), tree survival<br />

decreased (p < 0.038, r2= 0.36, n = 12), and tree susceptibility to b<strong>or</strong>er attack increased (p < 0.012, r2 = 0.45, n = 12).<br />

DISCUSSION<br />

Cannell et al. (1989) state that most tree species will benefit from an increase in annual mean temperature of I°C, but<br />

genetic increases will often cause growth reductions <strong>or</strong> m<strong>or</strong>tality. Similarly, Matyas (1994) states that an increase in temperature<br />

will positively affect tree growth only within each species' limits of physiological and ecological tolerance. Paper birch<br />

appears to have narrow limits of heat tolerance. In the present study, as the climatic differential increased f<strong>or</strong> individual seed<br />

sources, birch trees grew slower, experienced higher m<strong>or</strong>tality rates, and sustained higher rates of b<strong>or</strong>er attack. Such results<br />

strongly suggest that paper birch will experience dramatic decline along its southern range in response to climatic warming,<br />

and that the bronze birch b<strong>or</strong>er will likely be the principal m<strong>or</strong>tality agent involved in future birch decline. Similar predictions<br />

f<strong>or</strong> widespread decline of paper birch in response to global warming have been made by Past<strong>or</strong> and Post (1988),<br />

Overpeck et al. (1991), Botkin and Nisbet (1992), Reed and Desanker (1992), Reed et al. (1992b), Solomon and Bartlein<br />

(1992), and Jones et al. (1994). Several of the general circulation models predict reduced rainfall <strong>or</strong> increased rates of<br />

evapotranspiration in association with increased air temperatures (Karl et al. 1991, IPCC 1992). Thus, if higher temperatures<br />

are combined with reduced rainfall, then many tree species will likely be even m<strong>or</strong>e susceptible to attack by trunk-b<strong>or</strong>ing<br />

insects (Mattson and Haack 1987a, 1987b).<br />

As mentioned above, differences in sensitivity to photoperiod between n<strong>or</strong>thern and southern seed sources could<br />

partly explain the relatively po<strong>or</strong>er perf<strong>or</strong>mance exhibited by the m<strong>or</strong>e n<strong>or</strong>therly birch families in the present study. However,<br />

when I restricted the birch data set to seed sources that spanned as little as 1.5 ° and 2° latitude, the same patterns of<br />

reduced paper birch growth, survival, and resistance to bronze birch b<strong>or</strong>er persisted. In a Picea abies provenance study,<br />

Vaartaja (1959) rep<strong>or</strong>ted little difference in tree growth f<strong>or</strong> seed sources collected between 47°N and 52°N and then planted at<br />

47°N. Although the general rule f<strong>or</strong> photoperiod is that the further n<strong>or</strong>th the <strong>or</strong>igin, the m<strong>or</strong>e photoperiod sensitive the seed<br />

source, Beuker (1994) states that the exact effect of photoperiod cannot be predicted. Although changes in photoperiod are<br />

undoubtedly imp<strong>or</strong>tant, the significant relationships found here f<strong>or</strong> two relatively tight latitudinal bands, strongly suggest that<br />

the differences in seed source perf<strong>or</strong>mance in the present study were greatly influenced by temperature.<br />

Another possible reason f<strong>or</strong> the po<strong>or</strong>er growth of cold-adapted seed sources at warmer sites, is that they may have<br />

higher dark respiration rates than local sources (Kozlowski et al. 1991, Mebrahtu et al. 1991). Consequently, when growing<br />

in a warmer environment, cold-adapted sources expend m<strong>or</strong>e energy and have lower net assimilation rates than local sources.<br />

Reduced rates of net assimilation can help explain greater susceptibility to b<strong>or</strong>er attack since compromised trees would likely<br />

produce fewer defensive compounds and have reduced rates of diameter growth (Herms and Mattson 1992). Slow stem<br />

growth has been shown to increase susceptibility of paper birch to the bronze birch b<strong>or</strong>er (Herms 1991, Herms and Mattson<br />

1991).<br />

In Michigan, as was true f<strong>or</strong> much of N<strong>or</strong>th America, 1988 was a year of severe heat and drought (Kerr 1989, USGS<br />

1991). In the Great Lakes region, outbreaks of several cambial-feeding f<strong>or</strong>est insects occurred in 1988, including the bronze<br />

birch b<strong>or</strong>er (Haack and Mattson 1989, Jones et al. 1993). Part of the rapid increase in birch m<strong>or</strong>tality and b<strong>or</strong>er attack<br />

between years 12 (1987) and 15 (1990) in the present study (Figs. 3B, 3C) could have been triggered by the 1988 drought. In<br />

fact, it appears from the slopes of tree survival (Fig. 3B) and b<strong>or</strong>er attack (Fig. 3C) f<strong>or</strong> the four CD treatments, that the CD2 °,<br />

CD3 °, and CD4 ° trees were much m<strong>or</strong>e negatively impacted by the heat and drought than were the CD1 ° trees, suggesting<br />

that an increase in annual mean temperature of about 1°C will have limited detrimental effects on paper birch, but that<br />

increases of 2°C and m<strong>or</strong>e will be very damaging. In supp<strong>or</strong>t of <strong>this</strong> contention are studies in the Upper Peninsula of<br />

Michigan by Reed et al. (1992a) and Jones et al. (1993) that rep<strong>or</strong>t slower growth and higher m<strong>or</strong>tality rates f<strong>or</strong> paper birch,<br />

following the warm and dry years of 1987 and 1988 when mean summer temperatures were about 1°C above n<strong>or</strong>mal. In<br />

controlled studies on yellow birch, Betula alleghaniensis Britton, limited rootlet m<strong>or</strong>tality was observed with a I°C increase<br />

in soil temperature, but extensive rootlet m<strong>or</strong>tality occurred with increases of 6-7°C (Redmond 1955, 1957).<br />

241


As global warming occurs, I predict that those tree species with associated cambial-feeding insects that invade the<br />

trunk will be the first tree species to exhibit widespread m<strong>or</strong>tality. This guild of trunk-infesting, cambial-feeding insects<br />

poses the greatest lethal threat to their host trees (Haack and Slansky 1987, Mattson et al. 1988, Haack and Byler 1993).<br />

Trees are typically highly resistant to cambial-feeding insects, but during periods of stress, they become m<strong>or</strong>e susceptible to<br />

such insects (Mattson and Haack 1987a, 1987b; Millers et al. 1989). Some of the b<strong>or</strong>eal tree genera that occur in Michigan<br />

that can be killed by trunk-infesting, cambial-feeding insects, and thus will likely be m<strong>or</strong>e at risk as climatic warming occurs,<br />

include Abies, Betula, Larix, Picea, Pinus, Populus, and Tsuga (Table 3). Although climatic warming in the n<strong>or</strong>thern<br />

hemisphere is predicted to bring about tree m<strong>or</strong>tality along the southern edge of a species range, the n<strong>or</strong>thern range of these<br />

same tree species may expand (Cannell et al. 1989, Woodward 1992, Sykes and Prentice 1995). In fact, paper birch has been<br />

documented to be expanding its range n<strong>or</strong>thward into the tundra, possibly in response to climatic warming (Woodward 1992).<br />

Table 3.--B<strong>or</strong>eal tree genera of the Great Lakes region that will likely experience widespread insect-induced m<strong>or</strong>tality as a<br />

result of climatic warming and the c<strong>or</strong>responding trunk-infesting, cambial-feeding insect that will likely be the maj<strong>or</strong><br />

m<strong>or</strong>tality agent.<br />

Host tree Insect species<br />

Genus Common name Species Common name Family<br />

Abies Fir Pityokteins sparsus (LeConte) Balsam fir bark beetle Scolytidae<br />

Betula Birch Agrilus anxius G<strong>or</strong>y Bronze birch b<strong>or</strong>er Buprestidae<br />

Larix Larch Dendroctonus simplex LeConte Eastern larch beetle Scolytidae<br />

Picea Spruce Dendroctonus rufipennis (Kirby) Spruce beetle Scolytidae<br />

Pinus Pine Ips pini (Say) Pine engraver beetle Scolytidae<br />

Populus Aspen Agrilus liragus Barter and Brown Bronze poplar b<strong>or</strong>er Buprestidae<br />

Tsuga Hemlock Melanophilafulvoguttata (Harris) Hemlock b<strong>or</strong>er Buprestidae<br />

The present study shows that data from genetic test plantings of f<strong>or</strong>est tree species can provide insight into how treeinsect<br />

interactions may change as air temperatures increase in the future. Similar studies should be conducted by others.<br />

This would be particularly useful in areas where the local geography allows seed to be collected from several temperature<br />

regimes within a narrow latitudinal band and thereby allow the effects of temperature and photoperiod to be partly disentangled.<br />

ACKNOWLEDGMENTS<br />

I wish to thank the staff of Michigan State University, Fred Russ Experimental F<strong>or</strong>est, f<strong>or</strong> field assistance in <strong>this</strong><br />

study; Paul Bloese, Michigan State University, Department of F<strong>or</strong>estry and Michigan Cooperative Tree Improvement<br />

Program, f<strong>or</strong> making available early tree perf<strong>or</strong>mance data from <strong>this</strong> paper birch plantation; and Paul Bloese, Daniel Herms,<br />

Dow Gardens, and Robert Lawrence, <strong>USDA</strong> F<strong>or</strong>est Service, f<strong>or</strong> reviewing an earlier draft of <strong>this</strong> manuscript.<br />

LITERATURE CITED<br />

AKERS, R.C. and NIELSEN, D.G. 1990. Reproductive biology of the bronze birch b<strong>or</strong>er (Coleoptera: Buprestidae) on<br />

selected trees. J. Entomol. Sci: 25: 196-203.<br />

ANDERSON, R.F. 1944. The relation between host condition and attacks by the bronzed birch b<strong>or</strong>er. J. Econ. Entomol. 37:<br />

588-596.<br />

242


AUCLAIR, A.ND., WORREST, R.C., LACHANCE, D., and MARTIN, H.C. 1992. Climatic perturbations as a general<br />

mechanism of f<strong>or</strong>est dieback, p. 38-58. In Manion, P.D. and Lachance, D., eds. F<strong>or</strong>est Decline Concepts, American<br />

Phytopathological Society Press, St. Paul, Minnesota.<br />

AUCLAtR, A.N.D., LILL, J.T., and REVENGA, C. 1995. The role of climate variability and global warming in the dieback<br />

of n<strong>or</strong>thern hardwoods. Water, Air, and Soil Pollution (In press).<br />

AYRES, M.P. 1993. Plant defense, herbiv<strong>or</strong>y, and climate change, p. 75-94. In Kareiva, RM., Kingsolver, J.G., and Huey,<br />

R.B., eds. Biotic Interactions and Global Change. Sinauer Associates, Sunderland, Massachusetts.<br />

BALCH, R.E. and PREBBLE, J.S. 1940. The bronze birch b<strong>or</strong>er and its relation to the dying of birch in New Brunswick<br />

f<strong>or</strong>ests. F<strong>or</strong>. Chron. t6: 179-201.<br />

BALL, J. and SIMMONS, G. 1980. The relationship between bronze birch b<strong>or</strong>er and birch dieback. J. Arb<strong>or</strong>. 6:309-314.<br />

BARTER, G.W. 1957. Studies on the bronze birch b<strong>or</strong>er, Agrilus anxius G<strong>or</strong>y, in New Brunswick. Can. Entomol. 89: 12-<br />

36.<br />

BARTER, G.W. t 965. Survival and development of the bronze poplar b<strong>or</strong>er Agrilus liragus Barter and Brown (Coleoptera:<br />

Buprestidae). Can. Entomol. 997' 1063-1068.<br />

BENGTSSON, L. 1994. Climate change: Climate of the 21st century. Agric. F<strong>or</strong>. Mete<strong>or</strong>ol. 72: 3-29.<br />

BEUKER, E. 1994. Long-term effects of temperature on the wood production of Pinus sylvestris L. and Picea abies (L.)<br />

Karst. in old provenance experiments. Scand. J. F<strong>or</strong>. Res. 9: 34-45.<br />

BOTKIN, D.B. and NISBET, R.A. 1992. Projecting the effects of climate change on biological diversity in f<strong>or</strong>ests, p. 277-<br />

293. In Peters, R.L. and Lovejoy, T.E., eds. Global Warming and Biological Diversity. Yale Univ. Press, New Haven,<br />

CT.<br />

BRAATHE, R 1995. Birch dieback - caused by prolonged early spring thaws and subsequent frost. N<strong>or</strong>wegian J. Agric. Sci.<br />

Supple. No. 20: 1-59.<br />

CAMMELL, M.G. and KNIGHT, J.D. 1992. Effects of climatic change on the population dynamics of crop pests. Adv.<br />

Ecol. Res. 22: 117-162.<br />

CANNELL, M.G.R., GRACE, J., and BOOTH, A. 1989. Possible impacts of climatic warming on trees and f<strong>or</strong>ests in the<br />

United Kingdom: a review. F<strong>or</strong>estry 62: 337-364.<br />

CARLSON, R.W. and KNIGHT, F.B. 1969. Biology, taxonomy, and evolution of four sympatric Agrilus beetles b<strong>or</strong>er<br />

(Coleoptera: Buprestidae). Contrib. Amer. Entom_ol.Inst. 4: 1-105.<br />

CHITTENDEN, F.H. t898. A destructive b<strong>or</strong>er enemy of birch trees, with notes on related species. U.S. Dep. Agric. Div.<br />

Entomol. Bull (new series) 18:44-51.<br />

CLARK, J. and BARTER, G.W. 1958. Growth and climate in relation to dieback of yellow birch. F<strong>or</strong>. Sci. 4: 343-363.<br />

DEWAR, R.C. and WATT, A.D. 1992. Predicted changes in the synchrony of larval emergence and budburst under climatic<br />

warming. Oecologia 89: 557-559.<br />

EICHENLAUB, V.L., HARMAN, J.R., NURNBERGER, F.V.; and STOLLE, H.J. 1990. The Climatic Atlas of Michigan.<br />

University of Notre Dame Press, Notre Dame, Indiana. 165 p.<br />

FARMER, R.E., Jr. 1993. Latitudinal variation in height and phenology of balsam poplar. Silvae Genetica 42: 148-153.<br />

243


FLEMING, R.A. and VOLNEY, W.J.A. 1995. Effects of climate change on insect defoliat<strong>or</strong> population processes in<br />

Canada's b<strong>or</strong>eal f<strong>or</strong>est: Some plausible scenarios. Water Air Soil Pollut. 82: 445-454.<br />

FRANKLIN, J.F., SWANSON, F.J., HARMON, M.E., PERRY, D.A., SPIES, T.A., DALE, V.H., MCKEE, A., FERRELL,<br />

W.K., MEANS, J.E., GREGORY, S.V., LATTIN, J.D., SCHOWALTER, T.D., and LARSEN, D. 1992. Effects of<br />

global climatic change on f<strong>or</strong>ests in n<strong>or</strong>thwestern N<strong>or</strong>th America, p. 244-257. In Peters, R.L. and Lovejoy, T.E., eds.<br />

Global Warming and Biological Diversity. Yale Univ. Press, New Haven, CT.<br />

HAACK, R.A. 1992. Tree-stress-buprestid interactions: Our current understanding and future needs, p. 75-76. In Allen,<br />

D.C. and Abrahamson, L.E, eds. Proceedings: N<strong>or</strong>th American F<strong>or</strong>est Insect W<strong>or</strong>k Conference. 25-28 March 1991,<br />

Denver, CO. U.S. Dep. Agric. F<strong>or</strong>. Serv. Gen. Tech. Rep. PNW-294.<br />

HAACK, R.A. and ACCIAVATTI, R. E. 1992. Twolined chestnut b<strong>or</strong>er. U.S. Dep. Agric. F<strong>or</strong>. Serv. F<strong>or</strong>est Insect and<br />

Disease Leaflet I68. 12 p.<br />

HAACK, R.A. and BENJAMIN, D.M. 1982. The biology and ecology of the twolined chestnut b<strong>or</strong>er, Agrilus bilineatus<br />

(Coleoptera: Buprestidae), on oaks, Quercus spp., in Wisconsin. Can. Entomol. 114:385-396.<br />

HAACK, R.A. and BYLER, J.W. 1993. Insects and pathogens: regulat<strong>or</strong>s of f<strong>or</strong>est ecosystems. J. F<strong>or</strong>. 91 (9): 32-37.<br />

HAACK, R.A. and MATTSON, W.J. 1989. The long, hot summer of '88: They nibbled while the f<strong>or</strong>ests burned. Natural<br />

Hist<strong>or</strong>y (January) 56-57<br />

HAACK, R.A. and SLANSKY, F. 1987. Nutritional ecology of wood-feeding Coleoptera, Lepidoptera, and Hymenoptera, p.<br />

449-486. In Slansky, F. and Rodriguez, J.G., eds. Nutritional Ecology of Insects, Mites, and Spiders. John Wiley,<br />

New Y<strong>or</strong>k.<br />

HAWBOLDT, L.S. 1947. Aspects of yellow birch dieback in Nova Scotia. J. F<strong>or</strong>. 45: 414-422.<br />

HEDDEN, R.L. 1989. Global climate change: Implications f<strong>or</strong> silviculture and pest management, p. 555-562. In Proceedings,<br />

Fifth Biennial Southern Silvicultural <strong>Research</strong> Conference. U.S. Dep. Agric. F<strong>or</strong>. Serv. Gen. Tech. Rep. SO-74.<br />

HERMS, D.A. 1991. Variation in resource allocation patterns of paper birch: Evidence f<strong>or</strong> physiological tradeoffs among<br />

growth, reproduction and defense. Ph.D. Dissertation, Michigan State University, Department of Entomology, East<br />

Lansing, MI.<br />

HERMS, D.A. and MATTSON, W.J. 1991. Does reproduction compromise defense in woody plants? p. 35-46. In<br />

Baranchikov, Y.N., Mattson, W.J., Hain, F.R, and Payne, T.L., eds. F<strong>or</strong>est Insect Guilds: Patterns of Interaction with<br />

Host Trees. U.S. Dep. Agric. F<strong>or</strong>. Serv. Gen. Tech. Rep. NE-153.<br />

HERMS, D.A. and MATTSON, W.J. 1992. The dilemma of plants: To grow <strong>or</strong> defend. Quart. Rev. Biol. 67: 283-335.<br />

IPCC (Intergovernmental Panel on Climate Change). 1992. The Supplementary Rep<strong>or</strong>t to the IPCC Scientific Assessment.<br />

Houghton, J.T., Callander, B.A., and Varney, S.K., eds. Cambridge University Press, Cambridge, UK. 198 p.<br />

JONES, E.A., REED, D.D., MROZ, G.D., LIECHTY, H.O., and CATTELINO, RJ. 1993. Climate stress as a precurs<strong>or</strong> to<br />

f<strong>or</strong>est decline: paper birch in n<strong>or</strong>thern Michigan, 1985-1990. Can. J. F<strong>or</strong>. Res. 23: 229-233.<br />

JONES, E.A., REED, D.D., and DESANKER, RV. 1994. Ecological implications of projected climate change scenarios in<br />

f<strong>or</strong>est ecosystems of central N<strong>or</strong>th America. Agric. F<strong>or</strong>. Mete<strong>or</strong>ol. 72:31-46.<br />

KARL, T.R., HEIM, JR., R.H., and QUAYLE, R.G. 1991. The greenhouse effect in central N<strong>or</strong>th America: If not now,<br />

when? Science 251" 1058-1061.<br />

244 :_:<br />

i


KELLOMAKI, S., HANNINEN, H., and KOLSTROM, T. 1988. Model computations on the impacts of the climatic change<br />

on the productivity and silvicultural management of the f<strong>or</strong>est ecosystem. Silva Fennica 22: 293-305.<br />

KERR, R.A. 1989. 1988 ties f<strong>or</strong> the warmest year. Science 243 891.<br />

KOZLOWSKI, T.T., KRAMER, RJ., and PALLARDY, S.G. 1991. The Physiological Ecology of Woody Plants. Academic<br />

Press, San Diego. 657 p.<br />

LOERCH, C.R. and CAMERON, E.A. 1983. Determination of larval instars of the bronze birch b<strong>or</strong>er, Agrilus anxius<br />

(Coleoptera: Buprestidae). Ann. Entomol. Soc. Amer. 76: 948-952.<br />

LOERCH, C.R. and CAMERON, E.A. 1984. Within-tree distributions and seasonality of immature stages of the bronze<br />

birch b<strong>or</strong>er, Agrilus anxius (Coleoptera: Buprestidae). Can. Entomol. 116: 147-152.<br />

MACCRACKEN, M.C., BUDYKO, M.I., HECHT, A.D., and IZRAEL, Y.A. 1990. Prospects f<strong>or</strong> Future Climate: A Special<br />

US/USSR Rep<strong>or</strong>t on Climate and Climate Change. Lewis Publishers, Chelsea, MI. 270 p.<br />

MATTSON, W.J. and HAACK, R.A. 1987a The role of drought stress in provoking outbreaks of phytophagous insects, p.<br />

365-407. In Barbosa, P. and Schultz, J.C., eds. Insect Outbreaks: Ecological and Evolutionary Perspectives. Academic<br />

Press, Orlando.<br />

MATTSON, W.J. and HAACK, R.A. 1987b. The role of drought in outbreaks of plant-eating insects. BioScience 37:110-<br />

118<br />

MATTSON, W.J., LAWRENCE, R.K., HAACK, R.A., HERMS, D.A., and CHARLES, RJ. 1988. Plant defensive strategies<br />

f<strong>or</strong> different insect feeding guilds in relation to plant ecological strategies and intimacy of host association, p. 1-38.<br />

In Mattson, W.J., Levieux, J., and Bernard-Dagan, C. eds. Mechanisms of Plant Resistance to Insects. Springer-<br />

Verlag, New Y<strong>or</strong>k.<br />

MATYAS, C. 1994. Modeling climate change effects with provenance tests data. Tree Physiol. 14: 797-804.<br />

MDA (Michigan Department of Agriculture). 1971. Climate of Michigan by stations. Michigan Department of Agriculture,<br />

Climatological Division, Lansing, MI.<br />

MDA (Michigan Department of Agriculture). 1974. Supplement to the climate of Michigan by stations: mean temperature<br />

maps f<strong>or</strong> the period 1940-1969. Michigan Department of Agriculture, Climatological Division, Lansing, MI. 14 p.<br />

MEBRAHTU, T., HANOVER, J.W., LAYNE, D.R., and FLORE, J.A. 1991. Leaf temperature effects on net photosynthesis,<br />

dark respiration, and phot<strong>or</strong>espiration of seedlings of black locust families with contrasting growth rates. Can. J. F<strong>or</strong>.<br />

Res. 21" 1616-1621.<br />

MILLER, R.O., BLOESE, RD., HANOVER, J.W., and HAACK, R.A. 1991. Paper birch and European white birch vary in<br />

growth and resistance to bronze birch b<strong>or</strong>er. J. Amer. Soc. H<strong>or</strong>t. Sci. 116: 580-584.<br />

MILLERS, I., SHRINER, D.S., and RIZZO, D. 1989. Hist<strong>or</strong>y of hardwood decline in the eastern United States. U.S. Dep.<br />

Agric. F<strong>or</strong>. Serv. Gen. Tech. Rep. NE-126.<br />

NASH, R.W., DUDA, E.J., and GRAY, N.H. 1951. Studies on extensive dying, regeneration, and management of birch.<br />

Maine F<strong>or</strong>. Serv. Bull. 15.82 p.<br />

OVERPECK, J.T., BARTLEIN, RJ., and WEBB, III, T. 1991. Potential magnitude of future vegetation change in eastern<br />

N<strong>or</strong>th America: Comparisons with the past. Science 254: 692-695.<br />

PASTOR, J. and POST, W.M. 1988. Response of n<strong>or</strong>thern f<strong>or</strong>ests to CO2-induced climate change. Nature 334: 55-58.<br />

245


PETERS, R.L. 1992. Introduction, p. 3-114. In Peters, R.L. and Lovejoy, T.E., eds. Global Warming and Biological Diversity.<br />

Yale Univ. Press, New Haven, CT.<br />

PORTER, J.H., PARRY, M.L., and CARTER, T.R. 1991. The potential effects of climatic change on agricultural insect<br />

pests. Agric. F<strong>or</strong>. Mete<strong>or</strong>ol. 57: 221-240.<br />

REDMOND, D.R. 1955. Studies in f<strong>or</strong>est pathology. XV. Rootlets, myc<strong>or</strong>rhizae, and soil temperatures in relation to birch<br />

dieback. Can. J. Bot. 33:595-627.<br />

REDMOND, D.R. 1957. The future of birch from the viewpoint of diseases and insects. F<strong>or</strong>. Chron. 33: 25-30.<br />

REED, D.D. and DESANKER, RV. 1992. Ecological implications of projected climate change scenarios in f<strong>or</strong>est ecosystems<br />

in n<strong>or</strong>thern Michigan. Int. J. Biomete<strong>or</strong>ol. 36: 99-107.<br />

REED, D.D., JONES, E.A., HOLMES, M.J., and FULLER, L.G. 1992a. Modeling diameter growth in local populations: A<br />

case study involving four N<strong>or</strong>th American deciduous species. F<strong>or</strong>. Ecol. Mangt. 54: 95-114.<br />

REED, D.D., JONES, E.A., LIECHTY, H.O., MROZ, G.D. and JURGENSEN, M.E 1992b. Impacts of annual weather<br />

conditions on f<strong>or</strong>est productivity: A case study involving four N<strong>or</strong>th American deciduous f<strong>or</strong>est species. Int. J.<br />

Biomete<strong>or</strong>ol. 36:51-57.<br />

SCHLESINGER, M.E. and JIANG, X. 1991. Revised projection of future greenhouse warming. Nature 350: 219-221.<br />

SCHMIDTLING, R.C. 1994. Use of provenance tests to predict response to climatic change: loblolly pine and N<strong>or</strong>way<br />

spruce. Tree Physiol. 14: 805-817.<br />

SCRIBER, J.M. and GAGE, S.H. 1995. Pollution and global climate change: Plant ecotones, butterfly hybrid zones and<br />

changes in biodiversity, p. 319-344. In Scriber, J.M., Yoshitaka, T., and Lederhouse, R.C., eds. Swallowtail Ecology<br />

and Evolution. Scientific Publishers, Gainesville, FL.<br />

SHANDS, W.E. and HOFFMAN, J.S., eds. 1987. The Greenhouse Effect, Climate Change, and U.S. F<strong>or</strong>ests. The Conservation<br />

Foundation, Washington, D.C. 304 p.<br />

SLINGERLAND, M.V. 1906. The bronze birch b<strong>or</strong>er. C<strong>or</strong>nell Agric. Exp. Sta. Bull. 234: 63-78.<br />

SOLOMON, A.M. and BARTLEIN, RJ. 1992. Past and future climate change: Response by mixed-deciduous coniferous<br />

f<strong>or</strong>est ecosystems in n<strong>or</strong>thern Michigan. Can. J. F<strong>or</strong>. Res. 22: 1727-1738.<br />

SOLOMON, J.D. 1995. Guide to Insect B<strong>or</strong>ers in N<strong>or</strong>th American BroadleafTrees and Shrubs. U.S. Dep. Agric. F<strong>or</strong>. Serv.<br />

Agric. Handbk. AH-706. 735 p.<br />

SPAULDING, R and MACALONEY, H.J. 1931. A study of the <strong>or</strong>ganic fact<strong>or</strong>s concerned in the decadence of birch on cutover<br />

lands in n<strong>or</strong>thern New Englan d. J. F<strong>or</strong>. 29:1134-1149.<br />

SYKES, M.T. and PRENTICE, C. 1995. B<strong>or</strong>eal f<strong>or</strong>est futures: Modelling the controls on tree species range limits and<br />

transient responses to climate change. Water Air Soil Pollut. 82:415-428.<br />

SWAINE, J.M. 1918. A new f<strong>or</strong>est insect enemy of white birch. Can. J. F<strong>or</strong>. 14: 1928-1929.<br />

USGS (United States Geological Survey). 1991. National Water Summary 1988-1989 -- Hydrologic Events and Floods and<br />

Droughts. U.S. Geological Survey Water-Supply Paper 2375. 591 p.<br />

VAARTAJA, O. 1959, Evidence ofphotoperiodic ecotypes in trees. Ecol. Monogr. 29:91-111.<br />

246


WARGO, RM. and HAACK, R.A. 1991. Understanding the physiology of dieback and decline diseases and its management<br />

implications f<strong>or</strong> oak, p. 147-158. In The Oak Resource in the Upper Midwest: Implications f<strong>or</strong> Management Pro-<br />

ceedings, 3-6 June 1991, Winona, MN. Minnesota Extension Service, Univ. Minn. NR-BU-5663-S<br />

WILLIAMS, D.W. and LIEBHOLD, A.M. 1995. F<strong>or</strong>est defoliat<strong>or</strong>s and climate change: Potential changes in spatial distribu-<br />

tion of outbreaks of western spruce budw<strong>or</strong>m (Lepidoptera: T<strong>or</strong>tricidae) and gypsy moth (Lepidoptera:<br />

Lyrnantriidae). Environ. Entomol. 24: 1-9.<br />

WILSON, L. F. and HAACK, R.A. 1990. The bronze birch b<strong>or</strong>er. U.S. Dept. Agric. F<strong>or</strong>. Serv., N<strong>or</strong>th Central F<strong>or</strong>. Exp. Sta.<br />

N<strong>or</strong>thern Hardwood Note 7.09. 3 p.<br />

WOODWARD, EI. 1992. A review of the effects of climate on vegetation: Ranges, competition, and composition, p. 105-<br />

123. In Peters, R.L. and Lovejoy, T.E., eds. Global Warming and Biological Diversity. Yale Univ. Press, New Haven,<br />

CT.<br />

WRIGHT, J.W. 1976. Introduction to F<strong>or</strong>est Genetics. Academic Press, New Y<strong>or</strong>k.<br />

247


VARIATIONS IN SPRUCE NEEDLE CHEMISTRY AND IMPLICATIONS FOR<br />

THE LITTLE SPRUCE SAWFLY, PRISTIPHORA ABIETINA<br />

C. SCHAFELLNER, R. BERGER, J. MATTANOVICH, and E. FUHRER<br />

Institute of F<strong>or</strong>est Entomology, F<strong>or</strong>est Pathology, and F<strong>or</strong>est Protection<br />

Universit_it ftir Bodenkultur, Hasenauerstrasse 38, A 1190 Vienna, Austria<br />

INTRODUCTION<br />

The plant-herbiv<strong>or</strong>e system of N<strong>or</strong>way spruce, Picea abies Karst., and the little spruce Sawfly, Pristiph<strong>or</strong>a abietina<br />

Christ (Hymenoptera: Tenthredinidae), provides a good opp<strong>or</strong>tunity to study the synchronization between the host's bud burst<br />

and the beginning of larval feeding. Adults of the sawfly emerge in spring at about the same time as spruce bud break and<br />

oviposit in the soft, partially-expanded needles. Every single spruce bud is suitable f<strong>or</strong> oviposition f<strong>or</strong> only a few days. The<br />

emerging larvae almost always remain on the expanding bud where they hatch and only under starving conditions do they<br />

switch buds. After 2 <strong>or</strong> 3 weeks, depending on weather conditions, the larvae are fully grown and drop to the ground to f<strong>or</strong>m<br />

cocoons. They overwinter in the litter until the next spring.<br />

The larvae of the little spruce sawfly are early-spring feeders and restricted to feeding on expanding spruce needles.<br />

Their success and survival depend on the exact phenological coincidence of bud break and oviposition, because larvae can<br />

initiate feeding only in flushing buds. Seasonal variations in the timing of bud break can theref<strong>or</strong>e be of defensive value f<strong>or</strong><br />

the trees: trees that are genetically predisposed to flush needles earlier <strong>or</strong> later than the rest of the stand, thereby preceding <strong>or</strong><br />

following the sawfly swarming, may escape larval feeding.<br />

During the last 3 decades the sawfly has become an imp<strong>or</strong>tant spruce pest in several parts of Austria. The most<br />

severe attacks occur in lowland areas where the natural deciduous f<strong>or</strong>ests have been replaced by monocultures of N<strong>or</strong>way<br />

spruce. As the larvae feed exclusively on current-year needles, the affected trees do not die, but repeated infestations lead to<br />

def<strong>or</strong>med, bushy crowns and to reduced height and volume increments. In recent years, an increase in frequency and<br />

duration of epidemics has been observed, particularly in areas exposed to air pollutants like sulfur and nitrogen (Sierpinski<br />

1985, Berger 1992). There is good reason to believe that some air pollutants may fav<strong>or</strong> the sawfly's success, either by<br />

improving food quality due to changes in needle chemistry and/<strong>or</strong> by weakening the tree's defense system (Schafellner et al.<br />

1993, Schafellner et al. 1994, Berger und Katzensteiner 1994).<br />

This paper rep<strong>or</strong>ts a study on the most imp<strong>or</strong>tant food quality parameters of newly emerging needles during the sh<strong>or</strong>t<br />

period of larval feeding, and the changes that occur during the weeks of rapid growth. Additionally, the effect of excess<br />

nitrogen input (via fertilization) on needle chemistry and the nutritional value of the spruce needles f<strong>or</strong> the feeding larvae is<br />

demonstrated.<br />

METHODS<br />

Three even-aged stands of trees 16 years old, were selected f<strong>or</strong> study to see if food quality parameters show significant<br />

site-specific variation. Stand A is located at ca. 550 m elevation in Hochstrass, Lower Austria, and stands B and C are<br />

located at ca. 720 m elevation in the Hausruck, Upper Austria. Moderate sawfly attack was observed at stand A during the<br />

mid- 1980's. Throughout the 3 years of attack, tree flushing and sawfly attack f<strong>or</strong> every individual tree was rec<strong>or</strong>ded (Holzer<br />

t988). The stand was attacked irregularly so that affected and unaffected individuals often stood close together, although the<br />

trees did not vary in the timing of bud break and should have been suitable f<strong>or</strong> oviposition. From spring 1988 onwards, the<br />

sawflies vanished from the area so that during the 1989 sampling year, no frass was collected. Based on the data collected<br />

during the years of attack, 15 trees that had been affected once and 15 trees that remained unaffected, were selected f<strong>or</strong><br />

Mattson, W.J., Niemel_i, P., and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

248


investigation. Sampling was done on 3 occasions at 2-week-intervals during needle expansion in spring: needles were taken<br />

from twigs of the upper crown region, where the sawflies prefer to oviposit. To rec<strong>or</strong>d differences between the physiologically<br />

older needles at the bottom and the physiologically younger needles at the top of a given shoot, on the second and third<br />

occasion needles were separated acc<strong>or</strong>ding to their location on the expanding bud: top (t), middle (m), and bottom (b)<br />

needles.<br />

In contrast to stand A, the trees on stand B were undergoing an actual heavy sawfly attack during the sampling period<br />

in spring 1989. Additionally, large variation in the timing of bud burst of the trees was found: the early flushing individuals<br />

flushed 4 weeks bef<strong>or</strong>e the late flushing ones. Sawfly attack, however, occurred only on the late flushing trees. The early<br />

flushing trees (n = 10) were harvested 3 times (1,2,3), the late flushing trees (n = 10) only twice (la, 2a). After the second<br />

sampling of the attacked, late flushing trees (2a), the larvae were fully grown and were no longer feeding. Although the<br />

sampling on early and late flushing trees differed by m<strong>or</strong>e than 4 weeks, needles from sampling date 1 c<strong>or</strong>responded phenologically<br />

with 1a and needles from sampling date 2 c<strong>or</strong>responded with 2a.<br />

A fertilization experiment was done on stand C with late flushing, attacked trees. The stand was fertilized with 130<br />

kg nitrogen per ha/year, applied on 4 occasions: in summer 1989, in spring and summer 1990, and in spring 1991. Sampling<br />

was done during needle expansion in June, 1991, from 15 fertilized and 17 unfertilized (i.e., control) trees.<br />

Samples f<strong>or</strong> chemical analyses were put into liquid nitrogen, then freeze-dried and milled. Nitrogen (total and<br />

soluble) was determined by micro-Kjeldahl technique with selenium as catalyst. Soluble carbohydrates (sugars, cyclitols) as<br />

well as the <strong>or</strong>ganic acids (quinic acid, shikimic acid) were determined by GLC, starch content was estimated enzymatically as<br />

glucose-equivalents. Needle phenolics were defined and determined as the potential of the needle extract (done with boiling<br />

50% aqueous methanol) to precipitate bovine serum albumin (BSA). Abs<strong>or</strong>ption was measured at 590 nm. Fiber (cellulose,<br />

hemicellulose, lignin) was estimated after a method to determine the fiber content of cattle food: after hydrolyzation with<br />

strong acids and alkalines and heating up to 550 degrees, the remaining ash residue was weighed. All values are expressed on<br />

a dry weight basis.<br />

Larval experiments were perf<strong>or</strong>med at stand C with late flushing, fertilized, and unfertilized trees in spring 1991.<br />

Larvae from naturally attacked buds as well as larvae (lst to 3rd instar) placed on expanding buds and enclosed in cages were<br />

studied. After cocoon f<strong>or</strong>mation, larvae were collected, sexed, and weighed. Parasitized larvae were excluded from the<br />

analyses.<br />

RESULTS<br />

During initial feeding in mid-May (bud length 4 cm) the young sawfly larvae are confronted with both high total and<br />

soluble nitrogen (2% N tot. and 0.2% N sol.) and <strong>or</strong>ganic acid concentrations (12%). On the other hand, carbohydrates<br />

(sugars and cyclitols) and starch are at rather low levels (8% and 2%). Fiber was approximately 10% of the needle dry<br />

weight (dw). F<strong>or</strong> all these traits, damaged and undamaged trees did not differ (Fig. 1, upper graph). The crucial difference,<br />

however, was the phenolics: concentrations were significantly lower in the attacked trees than in the unattacked trees<br />

(p


40 ........................................ i.......................... T.......................... i........................... [........................... '..........=_ ......":..........................<br />

! i LDa,fe<br />

lii;iiiiiiii i i cte<br />

30 ............................. i ......................................................................................... ...................<br />

_' 20 .............................<br />

"0<br />

_ '!'!'!!!!!!!!!_!_!'!<br />

.............. ...... i<br />

10 I<br />

............<br />

.....<br />

i<br />

]iiiii!i_iiiiii]ii_!_l<br />

iiEii::iiiiiiiiii!i!i I<br />

....<br />

,_i_,_,_,E _<br />

! !ii[::::iiii_ii!i_i!_ ........<br />

, ii!iiiiii_i_iliiiiiii_i_l i!i!ii]i]iiii!]!iiii_;i!<br />

. ::ii_iiii_,i!i!i!i!i!i!i!<br />

0 i iiiii;iiii_;i!iii::i::iiil I ' _........... ::: :'_'_<br />

bud phenolics* total soluble soluble starch <strong>or</strong>ganic fiber<br />

expansion nitrogen nitrogen carbohydr, acids<br />

cm mg mg<br />

0 ...........................................................................................................................................................................................................................<br />

30 ............................................................................................................. ........................... _ ....................................... D affected<br />

_'20 .......................... _............. _.......................... _.......................... _........................... r.......................... i............................ i....<br />

l°una ecte°]<br />

Io OIl!iil ........ ................ .............<br />

bud expansion phenolics** total nitrogen soluble soluble starch <strong>or</strong>ganic acids fiber<br />

cm mg nitrogen mg carbohydrate<br />

Figure 1.--Spruce needle chemistry during initial (upper graph) and at the end (lower graph) of larval feeding on stand A<br />

(identical flushing time). Asterisks indicate significant differences between once affected (n = 15) and unaffected<br />

(n = 15)trees (levels of significance: *p < 0.05, ** p < 0.01).<br />

250


Figure 2.--Changes in protein precipitating ability ("phenolics") of spruce needle extracts during the period of larval<br />

feeding on stand A (identical flushing time). Asterisks indicate significant differences between once-affected (n<br />

= 115)and unaffected (n = 15) trees (levels of significance: * p < 0.05, ** p < 0.0l). Dates of sampling: I) 18<br />

May 1989, 2) 1 June 1989, and 3) 14 June 1989.<br />

500 I<br />

"_ 400 - I ........"<br />

_)<br />

= -i<br />

- _.......... - ____<br />

300 ..................................................................................<br />

I.....1.......................................................................<br />

O_<br />

"5 200-t-<br />

<<br />

co<br />

o !<br />

i<br />

1O0 ............ i<br />

0<br />

-i i<br />

1/la total 2/2a basal 2/2a top 3 basal 3 middle 3 top<br />

bud needles needles needles needles needles<br />

Figure 3.---Changes in protein precipitating ability ("phenolics") of"spruce needle extracts of early flushers (i.e., unaffected;<br />

n = 10), and late flushers (i.e., actually affected; n 10) during the period of needle expansion and larval<br />

feeding on stand B. Dates of sampling: early flushers: 1) 3 May 1989, 2) 17 May 1989, and 3) 1 June 1989; late<br />

flushers: la) 1 June 1989, and 2a) 14 June 1989.<br />

251


the control trees. Within the following 4 weeks, the phenolics decreased in the controls, but did not irathe attacked ones;<br />

concentrations being almost unchanged throughout larval feeding.<br />

The effect of N-fertilization on the expanding spruce needles was clear: total and soluble nitrogen were significantly<br />

higher in the fertilized trees and so were carbohydrates (p


q)<br />

J<br />

400 T J ! i i i i o control trees<br />

I _._1.................................................................................................................................................. [ I .........<br />

300 ................... ................. ....... i i i.................................................<br />

B<br />

ca i<br />

J D [] I I<br />

caE200-)_.........................................................................................................<br />

N...................................................................................................................<br />

_. m<br />

Q3 J I<br />

I<br />

} :=<br />

1,5 1,75 2 2,25 2,5 2,75 3 3,25 3,5<br />

% nitrogen dw<br />

Figure 5.mProtein precipitating ability Cphenolics") and nitrogen content of expanding spruce needles from individual<br />

fertilized (n = 15) and unfertilized (n = 17) trees. Date of sampling: 15 June 1991.<br />

I<br />

E<br />

-t []<br />

[]<br />

D control trees<br />

I<br />

[] fertilized trees []<br />

x:: [] !<br />

"_ 14 .... []<br />

12 ....................................................................................................<br />

_ ....................................................................... I<br />

c:_ [-3<br />

_-<br />

(D<br />

><br />

[]<br />

10 ....................................................... _ ......................... ] ............. m ..................... ...................................<br />

' i<br />

0 50 100 150 200 250 300<br />

mg BSA precipitated/g needle dw<br />

Figure 6._The relationship between the protein precipitating ability ("phenolics") of expanding spruce needles from<br />

individual fertilized (n = 9) and unfertilized (n = 8) trees, and the average weights of female larvae (n = 30-40 per<br />

tree) feeding on these trees.<br />

253


1992). However, several auth<strong>or</strong>s have rec<strong>or</strong>ded higher levels of phenolics in immature than in rnature leaves (Coley 1983,<br />

Puttick 1986, Mauffette and Oechel 1989, Hatcher 1990), as did the present study, but finally a gradual increase in concentration<br />

occurred with needle maturation (Schafellner et al. 1993). The decline of phenolics during the first month of needle<br />

expansion may be attributed to a dilution effect resulting from the addition of other leaf constituents and/<strong>or</strong> as phenolics are<br />

imp<strong>or</strong>tant precurs<strong>or</strong>s to lignin, to an inc<strong>or</strong>p<strong>or</strong>ation into the cell wall.<br />

Flushing needles of N<strong>or</strong>way spruce have the highest concentrations of key nutrients (total nitrogen, amino acids,<br />

water), however, they also contain the highest concentrations of secondary compounds (quinic acid, phenolics). Although<br />

nitrogen is an imp<strong>or</strong>tant fact<strong>or</strong> f<strong>or</strong> growth and survival and positive c<strong>or</strong>relations between needle carbohydrates and insect<br />

perf<strong>or</strong>mance have been demonstrated (Schopf 1986, Jensen 1988), we propose that between-tree differences in needle<br />

phenolics caused the observed pattern of sawfly attack. The amount of phenolics in the newly emerging needles may be<br />

crucial f<strong>or</strong> the feeding larvae. High phenolic levels suggest a potential defense mechanism of the host tree and seems to be<br />

determined m<strong>or</strong>e by intraspecific genetic differences than by site-dependent <strong>or</strong> seasonal fact<strong>or</strong>s (Lunderstfidt 1980). Additionally,<br />

we found no hint that needle quality deteri<strong>or</strong>ated f<strong>or</strong> the years after sawfly attack: nitrogen and carbohydrates did<br />

not decline and phenolics did not increase.<br />

Fertilization with N increased the nutritional value of spruce needles f<strong>or</strong> P. abietina larvae. This increase in nutritional<br />

value is based on an increase in the concentration of needle nitrogen and carbohydrates and a reduction in needle<br />

phenolics. Our results are consistent with several rep<strong>or</strong>ts that N fertilization increases the concentration of nitrogen and<br />

reduces the concentrations of carbon-based secondary compounds in leaves of woody plants (Waring et al. 1985, Haukioja et<br />

al. 1985, Glyphis and Puttick 1989) and that such changes in leaf chemistry are causally related to changes in leaf nutritional<br />

value f<strong>or</strong> immature insects (Bryant et al. 1987).<br />

An oversupply of N tends to change the metabolism of carbohydrates and nitrogen. Probably because of a higher<br />

demand f<strong>or</strong> carbohydrates during N-abs<strong>or</strong>ption and N-assimilation, the phenolic content of spruce needles decreases (Tuomi<br />

et al. 1990), whereas the concentration of total nitrogen and free amino acids increase with N fertilization <strong>or</strong> N deposition<br />

(N_isholm and Ericsson 1990). The low concentrations of phenolics in the leaves of fertilized trees may explain the increased<br />

insect susceptibility of trees exposed to N deposition and excess N (Berger and Katzensteiner 1994).<br />

In our study N fertilization also caused a significant increase in the needle concentrations of soluble carbohydrates<br />

(sugars, cyclitols), probably due to a stimulated growth of photosynthetic tissue, and a slight decrease in starch. N fertilization<br />

did not cause changes in beech foliar concentrations of total sugars, starch <strong>or</strong> soluble protein, but total phenolic compounds<br />

decreased by about a third (Balsber-Pahlsson 1992). Changes in total sugar content were found in the current-year<br />

needles of Scots pine (Ericsson 1979).<br />

Our results indicate that a shift in the carbon/nutrient balance of spruce by N fertilization increases the nutritional<br />

value of the needles f<strong>or</strong> P. abietina larvae in two ways: by increasing the concentration of nutritional chemicals (carbohydrates,<br />

nitrogen), and by reducing the concentrations of toxic and/<strong>or</strong> digestion-inhibiting chemicals (phenolics) in the needles.<br />

The experiment demonstrated that nitrogen fertilization enhanced larval growth. Low needle nitrogen concentrations and<br />

high concentrations of phenolics have been c<strong>or</strong>related with reduced larval weights of P. abietina feeding upon fertilized and<br />

unfertilized spruce trees. This supp<strong>or</strong>ts the idea that phenolics are specific agents f<strong>or</strong> protein deactivation.<br />

It has often been shown that the effect of air pollution may be fav<strong>or</strong>able to insect herbiv<strong>or</strong>es (Holopainen et al. 1991 ).<br />

Excess nitrogen supply (via atmospheric N-input) may promote the sawfly's success by improving its food quality: stimulated<br />

tree growth with high carbohydrate and nitrogen levels at the cost of limited allocation of resources f<strong>or</strong> the synthesis of<br />

defensive compounds (phenolics).<br />

SUMMARY<br />

The present study on the plant-herbiv<strong>or</strong>e system N<strong>or</strong>way spruce, Picea abies, and the little spruce sawfly, Pristiph<strong>or</strong>a<br />

abietina, illustrates variations in spruce needle chemistry as a potential fact<strong>or</strong> f<strong>or</strong> successful insect attacks. While nitrogen,<br />

carbohydrates, starch, <strong>or</strong>ganic acids, fiber, and water content of the new emerging needles showed no significant betweentree<br />

<strong>or</strong> site-specific differences during the time of larval feeding, the levels of the needle phenolics were significantly lower in<br />

254


attacked trees than in control trees. Successful larval feeding seems to be only possible on trees with low needle phenolics.<br />

High phenolic concentrations during early larval feeding suggest an effective defense mechanism of the host tree and seems<br />

to be genetically determined.<br />

A significant negative c<strong>or</strong>relation between needle nitrogen and phenolics in individual trees as well as the positive<br />

effect of nitrogen and the negative effect of phenolics on larval weight indicate a causal relationship between nitrogen and<br />

phenolics and larval growth.<br />

We also demonstrated the effects of excess atmospheric nitrogen input via N fertilization upon different chemical<br />

properties in spruce needles and the quality of these expanding needles as food f<strong>or</strong> the larvae. Nitrogen fertilization resulted<br />

in decreased phenolics, starch, <strong>or</strong>ganic acids, and fiber, increased nitrogen and carbohydrates, and in a better food value f<strong>or</strong><br />

the sawfly larvae. Larval weight was negatively c<strong>or</strong>related with needle phenolic concentrations. Our findings indicate that<br />

the quality of spruce needles as food f<strong>or</strong> the P abietina larvae are influenced in two ways: by increasing the concentrations<br />

of positive fact<strong>or</strong>s (e.g., nitrogen, carbohydrates) and decreasing the concentrations of negative fact<strong>or</strong>s (e.g., phenol ics, fiber,<br />

<strong>or</strong>ganic acids).<br />

ACKNOWLEDGEMENTS<br />

This study was funded by the Austrian Fonds of Scientific <strong>Research</strong>, P 7093-CHE: Excess Nitrogen in F<strong>or</strong>est<br />

Plant-Herbiv<strong>or</strong>e Associations. The auth<strong>or</strong>s wish to thank Ing. G. Mottlik, A. Stradner, and Mag. A. Straka f<strong>or</strong> technical<br />

assistance with the chemical analyses.<br />

LITERATURE CITED<br />

BALSBERG-PAHLSSON, A.M. 1992. Influence of nitrogen fertilization on minerals, carbohydrates, amino acids and<br />

phenolic compounds in beech (Fagus sylvatica L.) leaves. Tree Physiol. 10: 93-100.<br />

BERGER, R. 1992. Massenauftreten der Kleinen Fichtenblattwespe Pristiph<strong>or</strong>a abietina (Christ) im Hausruck. Teil I:<br />

Okologische Rahmenbedingungen in einem "untypischen" Befallsgebiet. Anz. Sch_idlingskde., Pflanzenschutz,<br />

Umweltschutz 65: 105-114.<br />

BERGER, R. and KATZENSTEINER, K. 1994. Massenauftreten der Kleinen Fichtenblattwespe Pristiph<strong>or</strong>a abietina<br />

(Christ) im Hausruck. Teil II: Immissions6kologische Einfltisse. J. appl. Ent. (in press).<br />

BRYANT, J.E, CHAPIN, F.S., KLEIN, D.R. 1983. Carbon/nutrient balance of b<strong>or</strong>eal plants in relation to vertebrate<br />

herbiv<strong>or</strong>y. Oikos 40: 357-368.<br />

BRYANT, J.E, CLAUSEN, T.E, REICHARDT, EB., McCARTHY, M.C., and WERNER, R.A. 1987. Effect of nitrogen<br />

fertilization upon the secondary chemistry and nutritional value of quaking aspen (Populus tremuloides Michx.)<br />

leaves f<strong>or</strong> the large aspen t<strong>or</strong>trix (Ch<strong>or</strong>istoneura conflictana (Walker)). Oecologia 73:513-517.<br />

COLEY, ED. 1983. Herbiv<strong>or</strong>y and defensive characteristics of tree species in a lowland tropical f<strong>or</strong>est. Ecol. Monogr. 53:<br />

209-233.<br />

ERICSSON, A. 1979. Effects of fertilization and irrigation on the seasonal changes of carbohydrate reserves in different<br />

age-classes of needle on 20-year-old Scots pine trees (Pinus sylvestris). Physiol. Plant. 45: 270-280.<br />

FEENY, EE 1970. Seasonal changes in oak leaf tannins and nutrients as a cause of spring feeding by winter moth caterpillars.<br />

Ecology 51:565-581.<br />

FEENY, EE 1976. Plant apparency and chemical defense. Recent Adv. Phytochem. 19: 1-40.<br />

255


GLYPHIS, J.R and PUTTICK, G.M. 1989. Phenolics, nutrition and insect herbiv<strong>or</strong>y in some garrigue and maquis plant<br />

species. Oecologia 78: 259-263.<br />

HATCHER, RE. 1990. Seasonal and age-related variation in the needle quality of five conifer species. Oecologia 85: 200-<br />

212.<br />

HAUKIOJA, E., NIEMELA, R, and SIREN, S. 1985. Foliage phenols and nitrogen in relation to growth, insect damage,<br />

and ability to recover after defoliation, in the mountain birch Betula pubescens ssp t<strong>or</strong>tuosa. Oecologia 65:214-222.<br />

HOLOPAINEN, J.K., KAINULAINEN, E., OKSANEN, J., WULFF, A., and KARENLAMPI, L. 1991. Effect of exposure<br />

to flu<strong>or</strong>ide, nitrogen compounds and SO 2on the numbers of spruce shoot aphids on N<strong>or</strong>way spruce seedlings.<br />

Oecologia 86: 51-56.<br />

HOLZER, K. and SCHULTZE, U. 1988. Die Bedeutung yon Herkunft und Austrieb bei einem Fichtenblattwespenbefall.<br />

Cbl. ges. F<strong>or</strong>stwesen 105:207-216.<br />

JENSEN, T.S. 1988. Variability of N<strong>or</strong>way spruce (Picea abies L.) needles; perf<strong>or</strong>mance of spruce sawflies (Gilpinia<br />

hercyniae Htg.). Oecologia 77:313-320.<br />

LUNDERST,ADT, J. 1980. Zur 6kophysiologischen Bedeutung yon Phenolen und Proteinen in Nadeln der Fichte (Picea<br />

abies Karst.). Z. Pflanzenern_ihr. Bodenkd. 143: 412-421.<br />

MATTSON, W.J. 1980. Herbiv<strong>or</strong>y in relation to plant nitrogen content. Ann. Rev. Ecol. Syst. 11: 119-161.<br />

MAUFFETTE, Y. and OECHEL, W.C. 1989. Seasonal variation in leaf chemistry of the coast live oak Quercus agrifolia<br />

and implications f<strong>or</strong> the Calif<strong>or</strong>nia oak moth Phryganidia calif<strong>or</strong>nica. Oecologia 79: 439-445.<br />

NASHOLM, T. and ERICSSON, A. 1990. Seasonal changes in amino acids, protein and total nitrogen in needles of fertilized<br />

Scots pine trees. Tree Physiol. 6: 267-281.<br />

PUTTICK, G.M. 1986. Utilization of evergreen and deciduous oaks by the Calif<strong>or</strong>nian oak moth, Phryganidia calif<strong>or</strong>nica.<br />

Oecologia 68: 589-594.<br />

SCHAFELLNER, C., BERGER, R., MATTANOVICH, J., and FOHRER, E. 1993. Pristiph<strong>or</strong>a abietina (Hym.,<br />

Tenthredinidae) - ein Bioindikat<strong>or</strong> ftir Luftverschmutzung? Besonderheiten unreifer Fichtennadeln als Larvenfutter.<br />

F<strong>or</strong>stw. Cb. 112: 116-128.<br />

SCHAFELLNER, C., BERGER, R., MATTANOVICH, J., and Ff0HRER, E. 1994. Food quality of spruce needles and the<br />

perf<strong>or</strong>mance of the Little Spruce Sawfly, Pristiph<strong>or</strong>a abietina Christ (Hym., Tenthredinidae). The protein precipitating<br />

ability of the young needles. Acta h<strong>or</strong>ticulturae (in press).<br />

SCHOPE R. 1986. The effect of secondary needle compounds on the development of phytophagous insects. F<strong>or</strong>. Ecol.<br />

Man. 15: 55-64.<br />

SCRIBER, J.M. and SLANSKY, E 1981. The nutritional ecology of immature insects. Ann. Rev. Entomol. 26: 183-211.<br />

SIERPINSKI, Z. 1985. Luftverunreinigungen und F<strong>or</strong>stsch_idlinge. J. appl. Ent. 99: 1-6.<br />

TUOMI, J., NIEMEL,_, R, and SIREN, S. 1990. The Panglossian paradigm and delayed inducible accumulation of foliar<br />

phenolics in mountain birch. Oikos 59: 399-410.<br />

WARING, R.H., McDONALD, A.J.S., LARSSON, S., ERICSSON, T., WIREN, A., ARWIDSSON, E., ERICSSON, A., and<br />

LOHAMMAR, T. 1985. Differences in chemical composition of plants grown at constant relative growth rates with<br />

stable mineral nutrition. Oecologia 65: 157-160.<br />

256


ACIDIC DEPOSITION, DROUGHT, AND INSECT HERBIVORY IN AN ARID<br />

ENVIRONMENT: ENCELIA FARINOSA AND TRIRHABDA GEMINATA IN<br />

SOUTHERN CALIFORNIA<br />

T.D. PAINE, R.A. REDAK, and J.T. TRUMBLE<br />

Department of Entomology, University of Calif<strong>or</strong>nia, Riverside, Calif<strong>or</strong>nia 92521, USA<br />

INTRODUCTION<br />

Except at the higher elevations, much of the arid southwestern United States and n<strong>or</strong>thern Mexico is not f<strong>or</strong>ested.<br />

The plant communities are dominated by widely spaced shrubs that are 1-3 m in height. Because these areas receive less than<br />

20 cm of annual precipitation and mean maximum summer temperatures of 38-43 °C, plant growth is often limited by the<br />

availability of moisture. Many of the species that are components of the Creosote Bush Scrub community are droughtdeciduous<br />

and remain d<strong>or</strong>mant during the dry periods (Munz and Keck 1968). Even in the slightly m<strong>or</strong>e moderate Coastal<br />

Sage Scrub community that receives up to 40 cm of annual precipitation and less extreme mean maximum summer temperatures,<br />

the plant species are adapted to extended periods without precipitation. Growth and flowering are often limited to<br />

those periods with available moisture.<br />

However, as the population of urban areas adjacent to these biotic communities grows, plants in those communities<br />

are subject to additional stresses arising from human activities. One of the most imp<strong>or</strong>tant anthropogenic stresses on plants is<br />

air pollution (Cowling 1982, Lee 1982, Linthurst et al. 1982, Treshow 1984, Shriner 1986). Although there are many<br />

different atmospheric pollutants capable of affecting plants, including ozone, NO x, SO, and PAN (peroxyacetyl nitrate), the<br />

impact on plants may be enhanced when the pollutants are combined with moisture. Wet acidic deposition in the f<strong>or</strong>m of<br />

acidic fog is of particular concern in southern Calif<strong>or</strong>nia. Acidic fog episodes in southern Calif<strong>or</strong>nia last from 2 to 12 hours<br />

and often have pH's of 2.0-3.0 (Waldman et al. 1982). Physical changes in leaf surface structure, lesions, leaching of foliar<br />

nutrients, decreased photosynthesis, and reduced growth and yield may result from long-term exposure to acidic fogs<br />

(Granett and Tayl<strong>or</strong> 1981, Granett and Musselman 1984, Musselman and Sterrett 1988, Musselman and McCool 1989,<br />

Paoletti et al. 1989, Takemoto et al. 1989, McCool and Musselman 1990, Mengel et al. 1990, Trumble and Walker 1991).<br />

Pollution-stressed plants also may be m<strong>or</strong>e susceptible to insects and diseases (Endress and Post 1985, Trumble et al. 1987,<br />

Trumble and Hare 1989, Jones and Coleman 1988, Paine et al. 1993). Consequently, if acidic deposition in the f<strong>or</strong>m of acidic<br />

fog occurs when the plants are actively growing and when the herbiv<strong>or</strong>es also are active, the pollution stress may have a<br />

significant impact on the fitness of the plant.<br />

We have studied the influence of acidic deposition on the interactions between brittle brush, Enceliafarinosa Gray<br />

(Asteraceae), a dominant shrub in the Creosote Bush Scrub and Coastal Sage Scrub communities of southern Calif<strong>or</strong>nia, and<br />

its primary herbiv<strong>or</strong>e, Trirhabda geminata H<strong>or</strong>n (Coleoptera: Chrysomelidae), a leaf beetle that feeds in both larval and adult<br />

stages on E. farinosa. The system is imp<strong>or</strong>tant because it is both widespread and an indicat<strong>or</strong> system f<strong>or</strong> damage that may be<br />

occuring in these fragile communities occupying harsh environments. The impact of pollution stresses on the interactions<br />

between plant and herbiv<strong>or</strong>e has been much less studied in desert than in f<strong>or</strong>est ecosystems (e.g., Johnson and Siccama 1983,<br />

Jones and Coleman 1988, Mengel et al. 1990); however, there may be much less environmental buffering in these desert<br />

communities f<strong>or</strong> additional stress imposed by anthropogenic encroachment.<br />

Acidic Deposition as a Single Stress<br />

Our previous studies have demonstrated significant effects of acidic deposition on the nutritional quality and palatability<br />

of E. farinosa to Z geminata in lab<strong>or</strong>at<strong>or</strong>y tests. Plants fogged three times f<strong>or</strong> 3 hours each time at pH 2.75 were<br />

compared to plants fogged with a control fog at pH 5.80. Acidic-fogged foliage had significantly higher concentrations of<br />

Mattson, W.J., Niemel_i, E, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

257


total nitrogen (2.33%) and higher soluble protein content (4.96 mg/g leaf tissue) when compared to leaves from controlfogged<br />

plants (1.94% total nitrogen and 3.94 mg/g leaf tissue soluble protein) (Paine et al. 1993). There were no differences<br />

between treatments in water content of the leaf tissues (74.8% f<strong>or</strong> acidic-fogged plants and 75% f<strong>or</strong> control-fogged plants).<br />

Both larvae and adult beetles preferred to feed on leaf tissue from acid-fogged plants (Paine et al. 1993). The studies<br />

demonstrated that the changes in Encelia foliage that alter insect consumption and growth indices can occur within 7 days of<br />

treatment with acid fogs.<br />

Acidic deposition has an indirect effect through E. farinosa upon the growth rate and biomass of T. geminata. Acidic<br />

fogging (pH 2.75 vs. a control fog of pH 5.8) <strong>or</strong>E. farinosa resulted in a 34% increase in average larval biomass gain<br />

(Control = 2.23 mg vs. Acid= 2.98 mg) and a 3 I% increase in larval growth rate (Control= 0.16 mg'mgl'd _vs.<br />

Acid=0.21 mg, mg_-d_). The effect of fogging on growth rate and biomass gain was strongest f<strong>or</strong> insects initiating feeding 14<br />

days after treatment applications. The effect of acidic fogging and the duration of the effect (days post treatment) were<br />

independent of each other (no interaction); the effect of acidic fogging upon growth rate and biomass gain was statistically<br />

constant through time. F<strong>or</strong> adult beetles, tissue consumption of acidic-fogged foliage was 65% greater than tissue consumption<br />

of control foliage (Control = 0.0964 cm2 vs. Acid = 0.16 cm2). F<strong>or</strong> larval beetles, tissue consumption of acidic-fogged<br />

foliage was over 400% greater than control foliage (Control = 0.057 cm2 vs. Acid = 0.259 cm 2) (Redak et al. In Prep.).<br />

Simultaneous Impact of Acidic Deposition and Drought as Multiple Stresses<br />

Drought stress has been shown to mitigate the impact of gaseous air pollutants and, to a certain extent, acidic wet<br />

deposition (acid rain and fog) on plant physiological and growth processes (Tingey and Hogsett 1985, Meier et al. 1990,<br />

Bender et at. 1991, Showman 199l, Beyers et al. 1992, Temple et al. 1992). Our field data investigating the impact of acidic<br />

fog upon the E. farinosa-Z geminata system are consistent with the the<strong>or</strong>y that drought stress may ameli<strong>or</strong>ate the impact of<br />

acidic fog upon the susceptibility of plants to insect herbiv<strong>or</strong>y (Warrington and Whittaker 1990, Von Sury and Fluckiger<br />

1991). Drought stress may make the plant less susceptible to the effects of the air pollution, <strong>or</strong> alternatively, the two stresses<br />

may act independently but antagonistically on the plant so that the net effect on herbiv<strong>or</strong>e success is neutral. Using an<br />

undeveloped area (remnant coastal sage-scrub habitat) on the UCR campus, we applied acidic (pH = 2.0) <strong>or</strong> control (pH =<br />

5.8) fog treatments (3, 3-hour exposures, every other day f<strong>or</strong> 5 days) to 40 mature E. farinosa plants (20 plants per treatment<br />

application). Following tog exposure, we caged 50 second instar T.geminata larvae in nylon screen bags on each treatment<br />

plant. Larvae were allowed to feed and grow on the plants f<strong>or</strong> 16 days at which time they were removed and their fresh and<br />

dry masses were determined. Additionally, 7 days after fog treatments, we conducted a 72-hour lab<strong>or</strong>at<strong>or</strong>y bioassay determining<br />

the consumption and growth rates of third instar T. geminata fed foliage taken from the field treated plants. After<br />

removal of experimental insects, foliage was collected from each plant and analyzed f<strong>or</strong> total N and percent water.<br />

Acidic fogging showed no effect on larval growth <strong>or</strong> consumption values <strong>or</strong> any plant foliage quality values (Table<br />

1). F<strong>or</strong> the field experiment described here, we feel the drought status of the experimental plants may account f<strong>or</strong> the lack of<br />

treatment effects (acidic fog effects) upon insect perf<strong>or</strong>mance. At the time of the experiment, southern Calif<strong>or</strong>nia was in the<br />

sixth to seventh year of a sustained drought. F<strong>or</strong> the year of <strong>this</strong> study, 1990, the UCR area had received less than 15%<br />

(approx. 40-50 ram) of n<strong>or</strong>mal precipitation by the end of the growing season (CIMIS weather data). Indeed approximately 1<br />

week after treatment applications, several experimental plants (of both treatments) began to show characteristic signs of<br />

drought-induced leaf senescence (leaf curling, browning, dropping older m<strong>or</strong>e mature leaves), and by the second week<br />

following removal of experimental insects, all of the plants were undergoing severe leaf-loss. Enceliafarinosa n<strong>or</strong>mally is a<br />

drought deciduous plant (Ehleringer and Clark 1988); however, in <strong>this</strong> case, the period of leaf-drop began approximately 6<br />

weeks to 2 months earlier than n<strong>or</strong>mal. Given the sustained 6 years of drought and the very early period of leaf abscission,<br />

we are confident that all experimental plants were experiencing relatively severe drought stress.<br />

The response of E. farinosa to drought stress includes (among other processes) an increase in leaf pubescence<br />

(Ehleringer and Bj<strong>or</strong>kman 1978, Ehleringer 1982). We suspect that (1) the experimental plants were so severely drought<br />

stressed that additional stress, in the f<strong>or</strong>m of acidic fog deposition, may have no further influence on the plants' suitability as<br />

a host f<strong>or</strong> 7: geminata and (2) the m<strong>or</strong>phological changes that occur in E. farinosa with drought stress are such that the plant<br />

may be physically protected to some extent from the impact of acidic fog (i.e., drought is functioning as a "filter" to the<br />

impact of acidic fog). The dense covering of pubescence (as a result of drought stress) could have physically shielded and/<strong>or</strong><br />

buffered the plant from acid tbg deposition (Cap<strong>or</strong>n and Hutchinson 1987, Musselman 1988).<br />

258


Table 1.--Impact of acidic fog upon the foliage quality of Enceliafarinosa and perf<strong>or</strong>mance of Trirhabda gerninata _._<br />

A. Field Bioassay.<br />

Final Insect Total Nitrogen Total Water<br />

Biomass (rag) Content (%) Content (%)3<br />

Fog pH 2.00 0.()049+0.0001 2.83+0.06 60.82+0.54<br />

Fog pH 5.80 0.0049_+0.0001 2.80+0.07 60.08+0.54<br />

B. Lab<strong>or</strong>at<strong>or</strong>y Bioassay.<br />

Insect Biomass Insect Growth Insect Consumption<br />

Gain (mg) Rate (rag.tugt.d: _) Rate (mg.mg_*d 1)<br />

Fog pH 2.00 0.0027+0.0002 0.9008+0.0816 27.729+4.171<br />

Fog pH 5.80 0.0026-+0.0003 0.8252+0.0826 26.179-+4.259<br />

There were no significant differences in any of the measured parameters between acid- and control-fogged treatments,<br />

ANOVA.<br />

2Values shown are means + 1 standard err<strong>or</strong>.<br />

3Analyses carried out on arcsine transf<strong>or</strong>med data. Values shown are back transf<strong>or</strong>med.<br />

CONCLUSIONS<br />

The implications of altered host plant quality f<strong>or</strong> insect herbiv<strong>or</strong>e development are considerable (Rhoades 1983,<br />

Strong et al. 1984, White 1984). <strong>Research</strong>ers have clearly demonstrated that nitrogen availability and plant secondary<br />

chemistry affect such basic insect life processes as growth rates, survival, and reproductive capacity (Onuf 1978, Rosenthal<br />

and Janzen 1979, Prestidge 11982,Prestidge and McNeill 1983). Since acidic fog incidents typically occur over broad<br />

geographical areas (Perkins 1974, Hoffman et al. 1985), the cumulative effects on insect herbiv<strong>or</strong>es at the population level<br />

may be quite significant. Thus, potential changes :in plant physiology due to stress <strong>or</strong> direct injury resulting from acidic fog<br />

may well have m<strong>or</strong>e serious consequences than previously believed. The dynamics of herbiv<strong>or</strong>ous insect populations<br />

associated with areas subjected to acidic fogs may be dramatically affected depending on the type and magnitude of the<br />

chemical changes that can occur in plants subjected to <strong>this</strong> type of pollution.<br />

By using E. farinosa and its most common natural herbiv<strong>or</strong>e, Z geminata, we are able to draw from and build upon a<br />

relatively th<strong>or</strong>ough understanding of the relationships between the effects of acidic fog on E. farinosa foliage quality and T.<br />

geminata success. Previous studies have shown that the maj<strong>or</strong> defensive compounds in E. farinosa are the sesquiterpene<br />

farinosin and the two chromenes euparin and encecalin (Wisdom and Rodriguez 1982, 1983; Wisdom et al. 1983; Wisdom<br />

1985, 1988). These compounds reduce growth rates, surviv<strong>or</strong>ship, and quantity of food consumed f<strong>or</strong> early instar larvae of T.<br />

geminata. Additionally the foliar concentrations of these compounds are highest at that time of year when early instars are<br />

present and actively feeding. We are currently investigating whether <strong>or</strong> not the concentrations of these compounds are<br />

affected by exposure to acidic fogs. Changes in the concentrations of the secondary chemistry of the plants may partially<br />

explain the results presented here. However, Wisdom (1985) showed that T. geminata larvae apparently feed on tissues high<br />

in total nitrogen content. We have shown that acidic fog exposure may change primary plant chemistry, resulting in increased<br />

plant total nitrogen and soluble protein. These increases are associated with increased larval preference (Paine et al. 1993)<br />

and may also be associated with better larval perf<strong>or</strong>mance. If the herbiv<strong>or</strong>e consumes m<strong>or</strong>e plant tissue as a direct consequence<br />

of an anthropogenic stress and decreases plant fitness, there may be a significant impact on the plant community if the<br />

stress continues f<strong>or</strong> a significant period of time. Although drought may ameli<strong>or</strong>ate the impact of the anthropogenic stress, the<br />

long-term consequences of the pollution stress on the ecosystem are not clear.<br />

259


LITERATURE CITED<br />

BENDER, J., TINGEY, D.T., JAGER, H.J., RODECAE K.D., and CLARK, C.S. 1991. Physiological and biochemical<br />

responses of bush bean (Phaseolus vulgaris) to ozone and drought stress. J. Plant Physiol. 137: 565-570.<br />

BEYERS, J.L., RIECHERS, G.H., and TEMPLE, P.J. 1992. Effects of long-term ozone exposure and drought on the<br />

photosynthetic capacity of ponderosa pine (Pinus ponderosa Laws.). New Phytol. 122:81-90.<br />

CAPORN, S.J.M. and HUTCHINSON, T.C. 1987. The influence of temperature, water and nutrient conditions during<br />

growth on the response of Brassica oleracea L. to a single, sh<strong>or</strong>t treatment with simulated acid rain. New Phytol.<br />

106:251-259.<br />

COWLING, E.B. 1982. A status rep<strong>or</strong>t on acid precipitation and its biological consequences as of April 1981. In D'itri, F.,<br />

ed. Acid Precipitation-Effects on Ecological Systems. Ann Arb<strong>or</strong> Press, Ann Arb<strong>or</strong>, MI.<br />

EHLERINGER, J.R. 1982. The influence of water stress and temperature on leaf pubescence development in Encelia<br />

farinosa. Am. J. Bot. 69: 670-675.<br />

EHLERINGER, J.R. and BJORKMAN, O. 1978. A comparison of photosynthetic characteristics of Encelia species possessing<br />

glabrous and pubescent leaves. Plant Physiol. 62: 185-190.<br />

EHLERINGER, J.R. and CLARK, C. 1988. Evolution and adaptation in Encelia (Asteraceae), p. 221-248. In Gottlieb, L.<br />

and Jain, S., eds. Plant Evolutionary Biology. Chapman and Hall, New Y<strong>or</strong>k.<br />

ENDRESS, A.G. and POST, S.L. 1985. Altered feeding preference of Mexican bean beetle Epilachna varivestis f<strong>or</strong><br />

ozonated soybean foliage. Environ. Pollut. 39: 9-16.<br />

GRANETT, A.L. and MUSSELMAN, R.C. 1984. Simulated acidic fog injureslettuce. Atmosph. Environ. 18: 887-890.<br />

GRANETT, A.L. and TAYLOR, O.C. I981. The effect of designated pollutants on plants. Fifth annual rep<strong>or</strong>t, AMRL-TR-<br />

80-112, USAF Aerospace Med. Res. Lab., Wright-Patterson AFB, Ohio.<br />

HOFFMAN, M.R., JACOB, D.J., WALDMAN, J. M., MUNGER, J.W., and FLAGAN, R.C. 1985. Characterization of<br />

reactants, reaction mechanisms, and reacting products in atmospheric water droplets: fog, cloud, dew and rain water<br />

chemistry. Final rep<strong>or</strong>t f<strong>or</strong> contractARB No. AZ-048-32. Calif<strong>or</strong>nia Air Resources Board.<br />

JOHNSON, A.H. and SICCAMA, T.G. 1983. Acid deposition and f<strong>or</strong>est decline. Environ. Sci. Technol. 17: 299A-304A.<br />

JONES, C.G. and COLEMAN, J.S. 1988. Plant stress and insect behavi<strong>or</strong>: Cottonwood, ozone and the feeding and oviposition<br />

preference of a beetle. Oecologia 76:51-56.<br />

LEE, J.J. 1982. The effects of acid precipitation on crops. In D'itri, E, ed. Acid precipitation-effects on ecological systems.<br />

Ann Arb<strong>or</strong> Press, Ann Arb<strong>or</strong>, MI.<br />

LINTHURST, R.A., BAKER, J.P., and BARTUSKA, A.M. 1982. Effects of acidic deposition: a brief review. In Frederick,<br />

E, ed. A Specialty Conference on Atmospheric Deposition. Air Pollution Control Association, Pennsylvania.<br />

MCCOOL, P.M. and MUSSELMAN, R.C. 1990. Injury of three <strong>or</strong>namental flower crops from simulated acidic fog. Plant<br />

Dis. 74: 310-312.<br />

MEIER, S., GRAND, L.E, SCHOENEBERGER, M.M., REINERT, R.A., and BRUCK, R.I. 1990. Growth, ectomyc<strong>or</strong>rhizae<br />

and nonstructural carbohydrates of loblolly pine seedlings exposed to ozone and soil water deficit. Environ. Poll. 64:<br />

11-27.<br />

260


MENGEL, K., BREININGER, T., and LUTZ, H.J. 1990. Effect of simulated acidic fog on carbohydrate leaching, CO 2<br />

assimilation and development of damage symptoms in young spruce trees (Picea abies L. Karst). Environ. and Exp.<br />

Bot. 30: 165-173.<br />

MUNZ, P.A. and KECK, D.D. 1968. A Calif<strong>or</strong>nia fl<strong>or</strong>a and supplement. University of Calif<strong>or</strong>nia Press, Los Angeles.<br />

MUSSELMAN, R.C. 1988. Acid neutralizing capacity of leave exposed to acidic fog. Environ. and Exp. Bot. 28: 27-32.<br />

MUSSELMAN, R.C. and MCCOOL, EM. 1989. Effects of acidic fog on productivity of celery and lettuce and impact on<br />

incidence and severity of diseases. Ann. Appl. Biol. 114: 559-565.<br />

MUSSELMAN, R.C. and J. L. STERRETT, J.L. 1988. Sensitivity of plants to acidic fog. J. Environ. Qual. I7: 329-333.<br />

ONUF, C.R 1978. Nutritive value as a fact<strong>or</strong> in plant-insect interactions with an emphasis on field studies. In Montgomery,<br />

G., ed. The Ecology of Arb<strong>or</strong>eal Foliv<strong>or</strong>es. Smithsonian Institution Press, Washington, DC.<br />

PAINE, T.D., REDAK, R.A., and TRUMBLE, J.T. 1993. Impact of acidic deposition on Enceliafarinosa Gray<br />

(Compositae: Asteraceae) and feeding preferences of Trirhabda geminata H<strong>or</strong>n (Coleoptera: Chrysomelidae). J.<br />

Chem. Ecol. 19:97-105<br />

PAOLETTI, E., GELLINI, R., and BARBOLANI, E. 1989. Effects of acid fog and detergents on foliar leaching of cations.<br />

Water, Air, and Soil Poll. 45: 49-61.<br />

PERKINS, H.C. 1974. Air Pollution. McGraw-Hill, New Y<strong>or</strong>k.<br />

PRESTIDGE, R.A. 1982. Instar duration, adult consumption, oviposition and nitrogen utilization efficiencies of leafhoppers<br />

feeding on different quality food (Auchen<strong>or</strong>rhyncha: Homoptera). Ecol. Entomol. 7: 91-101.<br />

PRESTIDGE, R.A. and MCNEILL, S. 1983. The imp<strong>or</strong>tance of nitrogen in the ecology of grassland Auchen<strong>or</strong>rhyncha. In<br />

Lee, J.A., McNeili, S., and R<strong>or</strong>ison, I.H., eds. Nitrogen as an Ecological Fact<strong>or</strong>. 22nd Symp. Brit. Ecol. Soc.<br />

Biackwell Scientific Publ., Oxf<strong>or</strong>d.<br />

REDAK, R.A., PAINE, T.D., and TRUMBLE, J.T. In prep. A determination of the direct impact of acidic fogging upon the<br />

success (m<strong>or</strong>tality and growth) of Trirhabda geminata (Coleoptera: Chrysomelidae). f<strong>or</strong> Oecologia.<br />

RHOADES, D.F. 1983. Herbiv<strong>or</strong>e population dynamics and plant chemistry. In Denno, R.F. and McClure, M.S., eds.<br />

Variable Plants and Herbiv<strong>or</strong>es in Natural and Managed Systems. Academic Press, London.<br />

ROSENTHAL, G.A. and JANZEN, D.H. 1979. Herbiv<strong>or</strong>es: Their Interaction with Secondary Plant Metabolites. Academic<br />

Press, London.<br />

SHOWMAN, R.E. 1991. A comparison of ozone injury to vegetation during moist and drought years. J. Air Waste Manage.<br />

Assoc. 41" 63-64.<br />

SHRINER, D.S. 1986. Terrestrial ecosystems: wet deposition. In Legge, A.H. and Rupa, S.V., eds. Air pollutants and their<br />

effects on the terrestrial ecosystem. Advances in Environmental Science and Technology, v 18. John Wiley and<br />

Sons, New Y<strong>or</strong>k.<br />

STRONG, D.R., LAWTON, J.H., and SOUTHWOOD, R. 1984. Insects on Plants. Harvard University Press, Cambridge,<br />

MA.<br />

TAKEMOTO, B.K., JOHNSON, A.G., PARADA, C.R., and OLSZYK, D.M. 1989. Physiology and yield of field-grown<br />

Brassica oleracea L. exposed to acidic fog. New Phytol. 112: 369-375.<br />

261


TEMPLE, RJ., RIECHERS, G.H., and MILLER, RR. 1992. Foliar injury responses of ponderosa pine seedlings to ozone,<br />

we and dry acidic deposition, and drought. Environ. and Exp. Bot. 32:101-113. '_<br />

TINGEY, D.T. and HOGSETT, W.E. 1985. Water stress reduces ozone injury via a stomatal mechanism. Plant Physiol. 77"<br />

944-947.<br />

TRESHOW, M. 1984. Air Pollution and Plant Life. John Wiley and Sons, New Y<strong>or</strong>k.<br />

TRUMBLE, J.T. and HARE, J.D. 1989. Acidic fog-induced changes in host-plant suitability. J. Chem. Ecol. 15: 2379-2390.<br />

TRUMBLE, J.T. and WALKER, G.E 1991. Acute effects of acidic fog on photosynthetic activity and m<strong>or</strong>phology of _<br />

Phaseotus lunatus. H<strong>or</strong>tScience 26:1531-1534.<br />

TRUMBLE, J.T., HARE, J.D., MUSSELMAN, R.C., and MCCOOL, E 1987. Ozone-induced changes in host-plant<br />

suitability: interactions of Keiferia lycopersicella and Lycopersicon esculentum. J. Chem. Ecol. 13:203-218.<br />

VON SURY, R. and FLUCKIGER, W. 1991. Effects of air pollution and water stress on leaf blight and twig cankers of<br />

London planes (Platanus X acerifolia (Ait.) Willd.) caused by Apiognomonia veneta (Sacc. and Speg.) Hohn. New<br />

Phytol. 118: 397-405.<br />

WALDMAN, J.M., MUNGER, J.W., JACOB, D.J., FLAGEN, R.C., MORGAN, J.J., and HOFFMAN, M.R. 1982. Chemical<br />

composition of acid fog. Science 218: 677-680.<br />

WARRINGTON, S. and WHITTAKER, J.B. 1990. Interactions between sitka spruce, the green spruce aphid, sulphur<br />

dioxide pollution and drought. Environ. Poll. 65: 363-370.<br />

WHITE, T.C.R. 1984. The abundance of invertebrate herbiv<strong>or</strong>es in relation to the availability of nitrogen in stressed food<br />

plants. Oecologia 63" 90-105.<br />

WISDOM, C. 1985. Use of chemical variation and predation as plant defenses by Enceliafarinosa against a specialist<br />

herbiv<strong>or</strong>e. J. Chem. Ecol. 11" 1553-1565.<br />

WISDOM, C. 1988. Comparisons of insect use and chemical defense patterns of two Son<strong>or</strong>an desert shrubs, p. 36-49. In<br />

Zahary, R.G., ed. Desert Ecology 1986: A <strong>Research</strong> Symposium. Southern Calif<strong>or</strong>nia Academy of Sciences and the<br />

Southern Calif<strong>or</strong>nia Desert Studies Cons<strong>or</strong>tium.<br />

WISDOM, C. and RODRIGUEZ, E. 1982. Quantitative variation of the sesquiterpene lactones and chromenes of Encelia<br />

farinosa. Biochem. Syst. and Ecol. 10: 43-48.<br />

WISDOM, C. and RODRIGUEZ, E. 1983. Seasonal age-specific measurements of the sesquiterpene lactones and<br />

chromenes of Enceliafarinosa. Biochem. Syst. Ecol. 11"43-48.<br />

WISDOM, C.S., SMILEY, J.T., and RODRIGUEZ, E. 1983. Toxicity and deterrency of sesquiterpene lactones and<br />

chromenes to the c<strong>or</strong>n earw<strong>or</strong>m (Lepidoptera: Noctuidae). J. Econ. Entomol. 76: 993-998.<br />

262


THE RESISTANCE OF SCOTCH PINE TO DEFOLIATORS<br />

V.I. GRIMALSKY<br />

<strong>Research</strong> Institute of F<strong>or</strong>estry, Gemol, Bel<strong>or</strong>ussia<br />

SUMMARY<br />

Scotch pine, Pinus sylvestris, is the most widely distributed tree species in Eurasia. It is often damaged by several<br />

defoliating insects: Dendrolimus pini, Panolisflammea, Bupalus piniarius, and Diprion pini. The foci of mass outbreaks of<br />

these insects typically occur in pure pine stands on po<strong>or</strong>, dry sandy soils, and on richer soils (loamy sands and sandy loams)<br />

depleted by long agricultural use.<br />

Chemical analyses of needles showed that N contents are always higher in pines growing outside (i.e., on richer,<br />

moister soils) than inside the foci of mass outbreaks. With respect to sugars, no differences were found. Larvae on trees<br />

enclosed in gauze bags, showed a high m<strong>or</strong>tality of early instars (I-III) in stands on rich, moist soils. Young larvae cause<br />

ole<strong>or</strong>esin droplets to fbrm where they chew into needles, the mechanism Ofpine resistance to them. Such exudations are<br />

weak from trees on po<strong>or</strong>, dry soils. Theref<strong>or</strong>e, larvae can feed on them without impediment, and their m<strong>or</strong>tality is low.<br />

To monit<strong>or</strong> <strong>this</strong> resistance, the ends of living needles (about 1/3 - 1/4 their length) on live trees, were cut. After 5<br />

minutes the clipped needles were sc<strong>or</strong>ed as follows: 0 - ole<strong>or</strong>esin does not exudate from the needle; 1 - a thin film of<br />

ole<strong>or</strong>esin appears on the cut, <strong>or</strong> separate tiny droplets which do not converge to a single lens; 2 - a small ole<strong>or</strong>esin lens<br />

appears (no m<strong>or</strong>e than 0.5-1.0 mm thick); 3 - a larger ole<strong>or</strong>esin lens appears. The following indices of July ole<strong>or</strong>esin exudation<br />

intensity (I) and efficiency (E) were calculated:<br />

I = (nl+2*n2+3*n3) / N E = (n2+n3) * 100 / N<br />

where I is a mean index of ole<strong>or</strong>esin exudation; nl, n2, n3 are number of needles with sc<strong>or</strong>es of 1, 2, <strong>or</strong> 3; N = total number<br />

of investigated needles (including those sc<strong>or</strong>ed as 0); and E = ole<strong>or</strong>esin exudation efficiency.<br />

The relationship between m<strong>or</strong>tality (M) of larvae (I-III instars) and (I) was tested using regression: M = a*I - b,<br />

where a and b are parameters specific f<strong>or</strong> each insect species. Thus, f<strong>or</strong> Dendrolimus pini, M = 65.0 1 - 34.4, f<strong>or</strong> Acantholyda<br />

stellata, M = 37.3 1- 8.3, f<strong>or</strong> Diprion pini, M = 66.9 1- 51.2, and f<strong>or</strong> Neodiprion sertifer, M = 42.4 1 - 29.9. The r2 values f<strong>or</strong><br />

these regressions vary from 0.53 - 0.79.<br />

One can calculate that Scotch pine is resistant to young larvae of Neodiprion sertifer beginning at I _>1.7 and E > 70.<br />

F<strong>or</strong> other defoliat<strong>or</strong>s, Scotch pine is resistant at I > 1.4 and E > 40. It should be noted that July ole<strong>or</strong>esin exudation intensity<br />

must be determined no later than 3-4 days after heavy rains when the humus layer moisture is > 6-8%, and the moisture of the<br />

lower layers (down to 50 cm depth) is _>3-4%, and air temperature, _>10 *C.<br />

Oie<strong>or</strong>esin exudation intensity from the needles of several pine species show that they can be divided into three<br />

groups: (1) low ole<strong>or</strong>esin exudation intensity (Pinus banksiana, P mugo, P maritima), (2) intermediate intensity (P<br />

sylvestris, P. pallasiana, R nigra, P strobus), <strong>or</strong> (3) high intensity (P. sibirica, P. cembra, P. pithyusa).<br />

The connection between the ole<strong>or</strong>esin exudation intensity of various pine species and their resistance to various<br />

defoliat<strong>or</strong>s is not perfect. The resistance of pine species depends also upon terpene composition of ole<strong>or</strong>esin.<br />

Mattson, W.J., NiemeRi, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

263


COMPANION PLANTING OF THE NITROGEN-FIXING GLIRICIDIA SEPIUM<br />

WITH THE TROPICAL TIMBER SPECIES MILICIA EXCELSA AND ITS IM-<br />

PACT ON THE GALL FORMING INSECT PHYTOLYMA LATA<br />

MICHAEL R. WAGNER 1, JOSEPH R, COBBINAH 2, and DANIEL A. OFORI z<br />

_School of F<strong>or</strong>estry, N<strong>or</strong>thern Arizona University, P.O. Box 15018, Flagstaff, Arizona 86011, USA<br />

2F<strong>or</strong>estry <strong>Research</strong> Institute of Ghana, UST P.O. 63, Kumasi, Ghana<br />

INTRODUCTION<br />

In Ghana, West Africa, there is increasing concern over the loss of valuable commercial timber species that contribute<br />

to the economic development of <strong>this</strong> country (Wagner and Cobbinah 1993). Part of the diminishing supply of timber relates<br />

to the high rates of def<strong>or</strong>estation (Zerbe et al. 1980, Mergen 1983, Repetto 1988). But an equally imp<strong>or</strong>tant fact<strong>or</strong> that is<br />

often overlooked is the failure of native species to regenerate following disturbance in the natural f<strong>or</strong>est <strong>or</strong> when planted in<br />

plantations (Fontaine 1985). F<strong>or</strong>est insects can significantly contribute to the po<strong>or</strong> regeneration rates of tropical timber<br />

species. An example of <strong>this</strong> is the complete failure to artificially regenerate Milicia excelsa, an extremely valuable tropical<br />

timber species in Africa, because of heavy damage by the gall-f<strong>or</strong>ming psyllid, Phytolyma lata (White 1968, Wagner et al.<br />

1991).<br />

Milicia excelsa and M. regia are imp<strong>or</strong>tant timber species in Africa, known in the timber trade as lroko, Mvule, and<br />

locally in Ghana as odum. The wood is durable, unif<strong>or</strong>mly grained, and extremely resistant to termites. While the species is<br />

a common component of the natural overst<strong>or</strong>y in much of the remaining tropical f<strong>or</strong>ests of West Africa, it is not currently<br />

regenerating and has a predicted resource life of less than 10 years in Ghana (Alder 1989). Ref<strong>or</strong>estation eff<strong>or</strong>ts have focused<br />

on establishing plantations of Imko which have failed in part because of massive damage by Phytolyma lata, the odum<br />

gallfly.<br />

The odum gallfly oviposits on new foliage of Milicia spp., and the first instar nymphs damage leaf tissue and induce<br />

galls (White 1968; Orr and Osei-Nkrumah 1978; Cobbinah 1986, 1988). The nymphs feed within the gall tissue. Numerous<br />

attack sites result in large gall masses. After about 2-3 weeks, the galls burst open releasing the adult psyllid. The gall tissue<br />

that has burst open exudes gall and leaf fluids that are heavily colonized by saprophytic fungi. These saprophytic fungi result<br />

in decay of leaf and stem tissue and dieback of young shoots (Wagner et al. 1991). Multiple attacks within the same season<br />

result in heavy damage and frequent death to young seedlings.<br />

Considering that conventional plantations had largely failed, we decided to examine the influence of companion<br />

planting of Gliricidia sepium, a well-known nitrogen-fixing agrof<strong>or</strong>estry species, and Milicia excelsa on odum gallfly attack.<br />

We hypothesized that Gliricidia sepium would provide nitrogen and overst<strong>or</strong>y shade m<strong>or</strong>e similar to conditions that occur in<br />

the natural tropical f<strong>or</strong>ests. In addition, we examined the influence of shading and nitrogen fertilization independently in an<br />

eff<strong>or</strong>t to determine which fact<strong>or</strong> contributed m<strong>or</strong>e to the observed effects. In <strong>this</strong> manuscript we rep<strong>or</strong>t that companion<br />

planting, shading, and fertilization all reduce attack and damage by the odum gallfly on Iroko.<br />

Mattson, W.J., Niemela, E, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

264


EXPERIMENTAL METHODS<br />

Experiments were carried out at the F<strong>or</strong>estry <strong>Research</strong> Institute's experimental nursery about 5 km south of the<br />

University of Science and Technology, Kumasi, Ghana. The area is within the moist semi-deciduous f<strong>or</strong>est zone (Hall and<br />

Swaine 1983), a common tropical f<strong>or</strong>est type in West Africa that generally supp<strong>or</strong>ts populations of Milicia excelsa. Seedlings<br />

used in the companion planting and shading experiments were collected from a single parent tree (AA 17) from the village of<br />

Abofour in the Ashanti Region of Ghana. Other studies have indicated that <strong>this</strong> particular provenance is of intermediate<br />

susceptibility to Phytolyma lata (Cobbinah and Wagner 1994). Seedlings used in the fertilizer trial were of mixed <strong>or</strong>igin.<br />

Seeds were collected in 1991, sown on raised beds, and transplanted into polyethylene bags. Trees were kept under shade<br />

and watered regularly until used in the experiment. The spacing on all experiments was 1m x 1m.<br />

Companion Planting Trial<br />

The companion planting trial was established in September 1991. The trial consisted of four rows of 10 (40 total)<br />

Milicia excelsa planted in alternating rows with Gliricidia sepium cuttings and four rows of 10 Milicia excelsa planted<br />

without G. sepium. At periodic intervals the following data were collected: seedling height, number of leaves, number of<br />

galls/plant. Regression analysis was used to estimate the number of galls in large gall masses as described by Cobbinah and<br />

Wagner (1994). F<strong>or</strong> <strong>this</strong> experiment, the data were averaged f<strong>or</strong> each row of 10 trees to reduce variability, consequently n = 4<br />

f<strong>or</strong> both treatments.<br />

Shading Trial<br />

The shading experiment was established in October 1992. The shade was created by supp<strong>or</strong>ting commercial shade<br />

cloth (Hummerts International) over wooden structures. Four levels of shade were established: full sun = no shade, light<br />

shade = 33% shade cloth, medium shade = 57% shade cloth, and deep shade = 82% shade cloth. Under each shade level, 50<br />

seedlings were planted in five rows of 10 seedlings each. Data collection was as indicated above. F<strong>or</strong> <strong>this</strong> experiment, n =<br />

50 per treatment.<br />

Fertilizer Trial<br />

The fertilizer experiment was established in December 1992 and ran f<strong>or</strong> 6 months. The effect of fertilizer was tested<br />

using a randomized complete block design with four replicates (25 seedlings/plot and 100 seedlings/block) and four treatments:<br />

control (no fertilizer), NPK ® (15-15-15), Phostrogen ® (10-4-22), Sampi® (8-3-3). Fertilizer was applied in the<br />

irrigation water at the rate recommended by the manufacturer. Seedlings were kept under shade. Data were collected as<br />

above and the average number of diebacks was determined.<br />

Statistical Analysis<br />

Data f<strong>or</strong> the companion planting trial were analyzed using a paired t-test on STATGRAPHICS software. Each row of<br />

10 trees was treated as the sample unit. Data f<strong>or</strong> the shading trial were transf<strong>or</strong>med [Logl0(x+l)] to meet assumptions of<br />

homogeneity of variance and were analyzed using an ANOVA on STATGRAPHICS software. Data f<strong>or</strong> fertilizer trials were<br />

analyzed using ANOVA and LSD multiple range test using STATGRAPHICS software. The P value to reject the null<br />

hypothesis was set at 0.05.<br />

RESULTS<br />

Companion Trial<br />

Milicia excelsa seedlings growing together with the nitrogen fixing Gliricidia sepium were not significantly taller<br />

(Fig. 1) and they did not produce m<strong>or</strong>e leaves (Fig. 2) than Milicia excelsa seedlings growing alone during any sample<br />

period. However, there were significantly fewer galls during two sample periods on Milicia grown with Gliricidia (Fig. 3).<br />

There was also a tendency f<strong>or</strong> fewer galls in the presence of Gliricidia f<strong>or</strong> most of the 1992 sample periods (Fig. 3).<br />

265


250<br />

With Gliricidia<br />

[] Without Gliricidia P=0,83<br />

200 ..................................<br />

P=0,81<br />

o 150 ............................... P--:o.28 '_.--0.4_,.<br />

.E P=0.21<br />

03<br />

©<br />

C"<br />

c 100<br />

aJ<br />

50<br />

0<br />

P=0.31<br />

P=0.29<br />

P=0.12<br />

P=0.32<br />

P=0.48<br />

1/92 2/92 3/92 4/92 9/92 11/92 1/93 3/93 5/93 7/93<br />

Sample Date<br />

Figure 1.--Influence of companion planting Gliricidia sepium with Milicia excelsa on M. excelsa growth at nine sample dates<br />

over 14 months. P-value based on paired t-test.<br />

C<br />

800 ................................................<br />

_- 600<br />

O<br />

With Gtiricidia<br />

[] Without Gliricidia<br />

(f} P=,).55 P=0.28<br />

> _n<br />

//A<br />

Z 400 ............................................. //A<br />

o ///1<br />

I,- P=0.82 //A<br />

//A<br />

//A<br />

E P=O.S6 //A<br />

L 200 P=0:66 .......................<br />

I'0-" .lo ///I<br />

//A<br />

P=0,19<br />

11/91 12/91 1/92 2/92 3/92 4/92 9/92 11/92 1/93<br />

Sample Date<br />

Figure 2.--Influence of companion planting Gliricidia sepium with Milicia excelsa on the number of" leaves of M. excelsa.<br />

P-value based on paired t-test.<br />

266


C<br />

w<br />

Q.<br />

1,200<br />

1,000<br />

o 800<br />

C_<br />

--- 600<br />

0<br />

[] With Gliricidia<br />

[] Without Gliricidia<br />

P=0.02<br />

C'} P=O.07<br />

IE 400 ............... P=o.s4 ///<br />

-1 /// ...... P=0.40 '<br />

"_ //I<br />

"+'" I1i<br />

1-0- 200 ............................ ///<br />

///<br />

///<br />

///<br />

P=0.96<br />

0 ///<br />

1/92 2/92 3/92 4/92 9/9211/92 1/93 3/93 5/93 7/93<br />

Sample Date<br />

Figure 3.--Influence of companion planting Gliricidia sepium with Milicia excelsa on the total number of Phytolyma lata<br />

galls on M. excelsa. P-value based on paired t-test.<br />

Shading Trial<br />

Shading had a significant positive impact on height (Fig. 4). After 1 year, trees grown at the highest level of shade<br />

(82%) were nearly 50% taller than trees at 57% shade and 100% taller than trees grown in full sun. In general, there were<br />

m<strong>or</strong>e leaves produced under shade than in full sunlight, but <strong>this</strong> pattern was not consistent across sample dates (Fig. 5).<br />

Despite the fact that m<strong>or</strong>e leaves were available to Phytolyma lata under shade, there were fewer galls per plant under shade<br />

(Fig. 6). While <strong>this</strong> pattern is statistically significant at only two of the sample dates, the pattern is very consistent across all<br />

sample dates.<br />

Fertilizer Trial<br />

Two of the fertilizer treatments, 15-15-15 and 10-4-22, significantly affected Milicia height when compared to the<br />

control (Fig. 7). In general the higher the nitrogen the greater the impact. The highest nitrogen content fertilizer also resulted<br />

in significantly fewer branch diebacks. Diebacks are positively related to the number of galls (Cobbinah and Wagner 1994).<br />

DISCUSSION<br />

We are unaware of other studies that have clearly linked companion planting <strong>or</strong> agrof<strong>or</strong>estry techniques and modifications<br />

in the incidence and impact of damaging insects in a tropical ecosystem. While the patterns we observed are not<br />

absolutely consistent f<strong>or</strong> all the sample periods, there is sufficient evidence to warrant further study.<br />

There is, however, a considerable body of evidence on the relationship between fertilization and insect attack (see<br />

Stark 1965, and Schowalter et al. 1986 f<strong>or</strong> reviews). Strauss (1987) observed that fertilization tended to have a positive<br />

impact on sap feeding insects and no effect on chewing insects. Several studies have examined the relationship between<br />

fertilization and population parameters of the gall-f<strong>or</strong>ming adelgid Adelges coolyei (Mitchell and Paul 1974, Johnson et aL<br />

1977, Mitchell and Miller 1976). In each case they found a general trendof fertilization increasing population parameters of<br />

Adelges coolyei. While these results are opposite to our own findings, the environmental conditions and plant species are<br />

267


7O<br />

Full sun P:O.00<br />

60 [] Light shade<br />

Medium shade<br />

50 []Deep shade<br />

E<br />

o<br />

.-, 40<br />

x: 30<br />

E -- P=0,29 -- P=0,06 P=0.0<br />

P=O.O0<br />

2O . _ -<br />

10 .. ...._<br />

0<br />

10/92 11/92 1/93 3/93 9/93<br />

Sample Date<br />

Figure 4.--Height (cm) of Milicia excelsa grown under four levels of shade. Light shade = 33%, Medium shade = 57%,<br />

Deep shade = 82%.<br />

00 ................................................................<br />

_-_Full sun P=O,O3<br />

[] Light shade ,_,_<br />

t_ [] Medium shade _\'_<br />

T__150 ................................ _\\, . . .<br />

_--° [] Deep shade :::.:_<br />

(1) P=O.06 : i """"<br />

__ 100 .......................................................<br />

..... \\\.<br />

0<br />

E ..... \\\\ t_.tt<br />

: ....<br />

c- 50 :::: ,-.-,-, ...... :ii<br />

:_\'h \\\\ :::: ;ii;_<br />

F_. : \\\\ i:ii i N_<br />

\\\\ :::::<br />

\\\\ :::<br />

:.: : \\',.\ ::,:<br />

,,;<br />

\\\\ :::; ....<br />

10/92 11/92 1/93 3/93<br />

Sample Date<br />

Figure 5.--Number of Milicia excelsa leaves produced under four levels of shade. Light shade = 33%, Medium shade --<br />

57%, Deep shade = 82%.<br />

268


0 ..............................................................<br />

Light shade P=0,05<br />

8 " _ Medium shade ............................... _ .....<br />

[] Deep shade<br />

l _._Control<br />

E 6 ...................................................... ........<br />

(_<br />

.. P=0.08 P=0,19<br />

--...... __ __<br />

"-_ P=0.03<br />

±<br />

0 --<br />

10/92 11/92 1/93 3/93 9/93<br />

Sample Date<br />

Figure 6.ENumber of Phytolyma tata galls per Milicia excelsa plant grown under four levels of shade. Light shade = 33%,<br />

Medium shade = 57%, Deep shade = 82%.<br />

60<br />

.................. _1_Height 15<br />

50 .....................<br />

b l_ Diebacks ............ b 'I<br />

09<br />

&"4o ..... -_<br />

o -<br />

v _O 03<br />

O3 "O<br />

•_ 30 cc-<br />

O {..-<br />

E _<br />

520 5 _x<br />

10<br />

0 0<br />

Control 15-15-15 10-4-22 8-3-3<br />

Fertilizer Treatment<br />

Figure 7.mMilicia excelsa height (cm) and average number of Phytolyma lata induced diebacks as the result of four fertilizer<br />

treatments.<br />

i<br />

269


very different from the Milicia/Phytolyma system we examined. There is very little consistency in the response of herbiv<strong>or</strong>es<br />

to fertilization of their hosts (Stark 1965, Schowalter et al. 1986). Insufficient evidence exists to postulate any general<br />

pattern in <strong>this</strong> regard.<br />

Hall and Swaine (1983) rep<strong>or</strong>ted that Iroko is not abundant in the evergreen tropical f<strong>or</strong>est because it requires high<br />

light f<strong>or</strong> germination and establishment. However, as previously noted, plantations of <strong>this</strong> species planted in full sun have<br />

been heavily attacked by the odum gaUfly and have largely failed. In contrast to the rep<strong>or</strong>t of Hall and Swaine (1983), we<br />

have been told by a tropical tree physiologist that the optimal photosynthesis f<strong>or</strong> Iroko seedlings occurs at 8-10% of full<br />

sunlight, suggesting <strong>this</strong> is a shade tolerant species (E.M. Veenendaal, pers. comm.). The results we rep<strong>or</strong>t in <strong>this</strong> paper<br />

indicate that the abundance and subsequent impact of <strong>this</strong> insect is greater in full sun than under shade.<br />

We recognize that there would be considerable utility in examining the relationship between Iroko and the odum<br />

gallfly under natural field conditions. However, naturally regenerating seedlings of <strong>this</strong> species are difficult to find. In a<br />

recent field excursion to the Tain II F<strong>or</strong>est Reserve, which is noted f<strong>or</strong> its high density of Iroko in the overst<strong>or</strong>y, we were able<br />

to locate only a single seedling in 12 person/hours of searching. The low incidence of Iroko seedlings has been confirmed<br />

anecdotally by several botanists and f<strong>or</strong>estry field surveys. Growing stock as of 1987 was immature stock (1-7ft girth)<br />

834,000m 3and mature stock (> 7ft girth) 5,269,000m 3 (Alder 1989).<br />

The results we rep<strong>or</strong>t suggest that standard agrof<strong>or</strong>estry practices such as mixed planting with nitrogen-fixing trees<br />

may have utility as part of a pest management program to control the odum gallfly. Other approaches including selection f<strong>or</strong><br />

genetic resistances have resulted in considerable success (Cobbinah and Wagner 1994). Cooperative eff<strong>or</strong>t between silviculturists,<br />

physiologists, and genetists is required to fully understand how to reduce the impact of <strong>this</strong> insect and provide f<strong>or</strong> a<br />

continuing supply of one of Africa's most valuable timber species.<br />

SUMMARY<br />

Planting the nitrogen-fixing Gliricidia sepium in conjunction with the tropical timber species Milicia excelsa reduces<br />

the abundance and damage caused by the gall-f<strong>or</strong>ming pysUid, Phytolyma lata. This effect appears to be related to both the<br />

increased shading and enhanced nitrogen environment. Shading independently has a strong impact on gall abundance with<br />

the lowest gall abundance occurring under the highest level of shade (82%). Gall abundance is reduced under shade even<br />

though m<strong>or</strong>e leaves are available to the insect. Fertilization at the highest levels of nitrogen also reduces the abundance of<br />

galls.<br />

ACKNOWLEDGEMENT<br />

These studies were supp<strong>or</strong>ted by the International Tropical Timber Organization through research grant No. PD75/90<br />

to J.R. Cobbinah and M.R. Wagner.<br />

LITERATURE CITED<br />

ALDER, D. 1989. Naturalf<strong>or</strong>est increment, growth and yield, p. 47-52. In Ghana F<strong>or</strong>est Invent<strong>or</strong>y Proceedings. Overseas<br />

Dev. Agency/Ghana F<strong>or</strong>. Dep., Accra.<br />

COBBINAH, J.R. t986. Fact<strong>or</strong>s affecting the distribution and abundance of Phytolyma lata (Homoptera:PsyUidae). Insect<br />

Sci. Appl. 7: 1111-115.<br />

COBBINAH, J.R. 1988. The biology, seasonal activity and control of Phytolyma lata. IUFRO Regional W<strong>or</strong>kshop on Pests<br />

and Diseases of F<strong>or</strong>est Plantations, June 5-10, 1988, Bangkok, Thailand.<br />

COBBINAH, J.R. and WAGNER, W.R. 1994. Phenotypic variation in Milicia excelsa toattack by Phytolyma lata.<br />

Biotropica. (in review).<br />

270


FONTAtNE, R.G. 1985. F<strong>or</strong>ty years of f<strong>or</strong>estry at FAO: some personal reflections. Unasylva. 37(148): 5-14.<br />

HALL, J.B. and SWAINE, M.D. 1981. Distribution and Ecology of Vascular Plants in a Tropical Rain F<strong>or</strong>est: F<strong>or</strong>est<br />

Vegetation of Ghana. W. Junk Publishers, The Hague. 383 p.<br />

JOHNSON, J.M., LITTLE, M.R, and GARA, R.I. 1977. Effects of different N-fertilizers and concentrations on Adelges<br />

cooleyi populations. J. Econ. Entomol. 70(4): 527-528.<br />

MERGEN, F. 1983. Tropical F<strong>or</strong>estry: A Challenge to the Profession of F<strong>or</strong>estry. Yale University Printing Service, New<br />

Haven, CT. 23 p.<br />

MITCHELL, R.G. and PAUL, H.G. 1974. Field fertilization of Douglas-fir and its effects on Adelges cooleyi populations.<br />

Environ. Entomol. 3: 501-504.<br />

MITCHELL, R.G. and MILLER, R.E. 1976. Effects of Douglas.fir fertilization on the Cooley spruce gall aphid. Am.<br />

Christ. Tree J. 20(1): 29.<br />

ORR, R.L. and OSEI-NKRUMAH, A. 1978. Progress Rep<strong>or</strong>t on Phytolyma lata (Hemiptera:Psyllidae). F<strong>or</strong>est Products<br />

<strong>Research</strong> Institute. Kumasi, Ghana.<br />

REPETTO, R. 1988. The F<strong>or</strong>est f<strong>or</strong> the Trees? Government Policies and the Misuse of F<strong>or</strong>est Resources. W<strong>or</strong>ld Resources<br />

Institute. Washington, DC. 105 p.<br />

SCHOWALTER, T.D., HARGROVE, W.W., and CROSSLEY, D.A. 1986. Herbiv<strong>or</strong>y in f<strong>or</strong>ested ecosystems. Ann. Rev.<br />

Entomol. 31: 177-196.<br />

STARK, R.W. 1965. Recent trends in f<strong>or</strong>est entomology. Ann. Rev. Entomol. 10: 303-324.<br />

STRAUSS, S.Y. 1987. Direct and indirect effects of host-plant fertilization on an insect community. Ecology 68: 1670-<br />

1678.<br />

WAGNER, M.R., ATUAHENE, S.K.N., and COBBINAH, J.R. 1991. F<strong>or</strong>est Entomology in West Tropical Africa: F<strong>or</strong>est<br />

Insects of Ghana. Kluwer Academic Publishers, D<strong>or</strong>drecht. 210 p.<br />

WAGNER, M.R. and COBBINAH, J.R. 1993. Def<strong>or</strong>estation and sustainability in Ghana: the role of tropical f<strong>or</strong>ests. J. F<strong>or</strong>.<br />

91(6): 35-39.<br />

WHITE, M.G. 1968. <strong>Research</strong> in Nigeria on the Iroko Gall Bug (Phytolyma spp.). Nigeria F<strong>or</strong>. Inf. Bull. 18. 72 p.<br />

ZERBE, J.J., WHITMORE, J.L., WAHLGREN, H.E., LAUNDRIE, J.F., and CHRISTOPHERSEN, K.A. 1980. F<strong>or</strong>estry<br />

activities and def<strong>or</strong>estation problems in developing countries. Rep<strong>or</strong>t to Office of Science and Technology. U.S.<br />

Department of Agriculture, F<strong>or</strong>est Service, F<strong>or</strong>est Products Lab<strong>or</strong>at<strong>or</strong>y. 115 p.<br />

271


FUNCTIONAL HETEROGENEITY OF FOREST LANDSCAPES:<br />

HOW HOST DEFENSES INFLUENCE EPIDEMIOLOGY<br />

OF THE SOUTHERN PINE BEETLE<br />

ROBERT N. COULSON, JEFFREY W. FITZGERALD, BRYAN A. MCFADDEN,<br />

PAUL E. PULLEY, CLARK N. LOVELADY, and JOHN R. GIARDINO<br />

Knowledge Engineering Lab<strong>or</strong>at<strong>or</strong>y, Department of Entomology and Department of Geography, Texas A&M University<br />

College <strong>Station</strong>, Texas 77843, USA<br />

INTRODUCTION<br />

Insect outbreaks are autogenic disturbances that are n<strong>or</strong>mally observed at the landscape scale as levels of herbiv<strong>or</strong>y<br />

above an average <strong>or</strong> expected amount. When fav<strong>or</strong>able environmental conditions coincide with optimal resource availability,<br />

populations increase in size and outbreaks occur (Rykiel et al. 1988). Epidemiology is the study of causes f<strong>or</strong> change in the<br />

distribution and abundance of an insect species at tile landscape level of ecological integration. Typically, studies of epidemiology<br />

have considered herbiv<strong>or</strong>e populations in the context of their impact on food and habitat resources. In <strong>this</strong> paper we<br />

focus on the association of f<strong>or</strong>est landscape structure and the process of herbiv<strong>or</strong>y, i.e., the effect of landscape pattern on the<br />

process of herbiv<strong>or</strong>y (Turner 1989). Our goal is to examine epidemiology of the southern pine beetle, Dendroctonusfi'ontalis<br />

Zimmermann (Coleoptera: Scolytidae), in the context of functional heterogeneity of f<strong>or</strong>est landscapes. Our objectives are:<br />

(i) to review briefly the current state of knowledge on epidemiology of D. frontalis, (ii) to consider the concept of functional<br />

heterogeneity of f<strong>or</strong>est landscapes and how it relates to epidemiology of D. j?ontalis, (iii) to consider ways and means of<br />

measuring functional heterogeneity of f<strong>or</strong>est landscapes, and (iv) to illustrate the relation of functional heterogeneity and<br />

herbiv<strong>or</strong>y by the insect.<br />

Although emphasis is directed to D. frontalis, it is notew<strong>or</strong>thy that <strong>this</strong> insect is a member of a guild consisting of<br />

four other bark beetle species: D. terebrans (Olivier), Ips calligraphus (Germ_), L grandicollis (Eichhofl), and I. avulsus<br />

(Eichhoff) (Coleoptera: Scolytidae). The population systems of these insects are interrelated. How the members contribute<br />

to epidemiology of the guild is unknown.<br />

Background<br />

Epidemioiogy of the Southern Pine Beetle<br />

[=lammet al. (1988) provide a review of literature on the natural hist<strong>or</strong>y of D. frontalis. A general conceptual model<br />

of epidemiology of the insect was proposed by Coulson et aL (1983, 1985a, b, 1986a) and Rykiel et al. (1988). This model<br />

considers the process of herbiv<strong>or</strong>y in the context of variables thought to influence epidemiology, i.e., mete<strong>or</strong>ological conditions,<br />

the lighming disturbance regime, f<strong>or</strong>est stand conditions, background levels of the insect, and landscape structure.<br />

Studies by Flamm and Coulson (1988), Flamm et al. (1992), Lovelady et al. (199 I), and Lovelady (1994) have employed<br />

empirical methods to address specific aspects of the conceptual model.<br />

Outbreaks of D. frontalis are the result of population fluctuations observed at the landscape scale. There are three<br />

separate processes associated with D. fi'omalis epidemiology. The first, which operates at the population and community<br />

levels of integration, involves fact<strong>or</strong>s associated with initiation of infestations. The second, which operates at the ecosystem<br />

level of integration, deals with fact<strong>or</strong>s that influence growth of established infestations. The third, which operates at the<br />

landscape level of integration, deals with the netw<strong>or</strong>k of habitat units and population centers occurring in a heterogeneous<br />

land area consisting of multiple interacting ecosystems. An insect outbreak is an epidemiological condition that represents a<br />

composite of the three processes operating across different scales of time and space. The actual process of herbiv<strong>or</strong>y by D.<br />

fromalis occurs at the population and community levels of ecological integration and results in m<strong>or</strong>tality to host pines.<br />

However, the effects are propagated through component ecosystems to the landscape level (Coulson et al. 1986b, Rykiel et<br />

al. 1988, Lovelady et al. 1991).<br />

Mattson, W.J., Niemel_, R, and Rousi, M., eds. 1996. Dynamics of f<strong>or</strong>est herbiv<strong>or</strong>y: quest f<strong>or</strong> pattern and principle. <strong>USDA</strong><br />

F<strong>or</strong>. Serv. Gen. Tech. Rep. NC-183, N.C. F<strong>or</strong>. Exp. Sta., St. Paul, MN 55108.<br />

272


Host Defenses<br />

An imp<strong>or</strong>tant component of epidemiology of bark beetles centers on the interaction of the insect with its host.<br />

Considerable eff<strong>or</strong>t has been directed to research on host defense mechanisms, because of their critical role in regulating<br />

colonization by bark beetles. In pines (Pinus spp.) both pref<strong>or</strong>med and induced defense systems have been identified.<br />

During colonization, bark beetles directly confront the defense mechanisms of the host. The colonization process is greatly<br />

influenced by variation in defense capacity associated with different pines, (e.g., Iongleaf pine is m<strong>or</strong>e resistant to D. frontalis<br />

colonization than loblolly <strong>or</strong> sh<strong>or</strong>tleaf), and with different individuals of the same host species. Within a f<strong>or</strong>est landscape<br />

there is also a seasonal variation in host defense capability. L<strong>or</strong>io (I988) suggested that n<strong>or</strong>mal physiological changes<br />

associated with tree phenological development and response to variable environmental conditions cause regular and predictable<br />

fluctuations in resistance/susceptibility to D. frontalis. Lovelady (1994) provides a comprehensive review and interpretation<br />

of seasonal variation in host defense in relation to population dynamics ofD. frontalis.<br />

Lightning-struck hosts, which have greatly reduced capacity f<strong>or</strong> defense against colonization, serve as refuges f<strong>or</strong><br />

bark beetles. F<strong>or</strong> many years f<strong>or</strong>esters have recognized that infestations olD. frontalis are often associated with a lightningstruck<br />

host. The initial conceptual model of epiderniology of D. frontalis included lightning-struck hosts as a prominent<br />

component, although only circumstantial evidence supp<strong>or</strong>ted the contention (Coulson et al. 1983). Subsequent research has<br />

suggested that lightning-struck hosts are an essential feature of the natural hist<strong>or</strong>y of the insect. This research considered the<br />

interaction of D. frontalis with disturbed hosts (Coulson et al. 1986a, Flamm and Coulson 1988, Flamm et al. 1992) as well<br />

as an analysis of lightning as a disturbance regime (Lovelady et al. 1991, Lovelady 1994).<br />

Of particular concern in the study of epidemiology is how host defenses are distributed at the landscape scale.<br />

Herbiv<strong>or</strong>y by D. frontalis produces disturbance patches (infestations) in the f<strong>or</strong>est matrix that can be observed relative to<br />

other features of the landscape, i.e., the crowns of infested trees become discol<strong>or</strong>ed and are visually detectable. The disturbance<br />

patches are typically associated with old-growth f<strong>or</strong>est stands occurring on po<strong>or</strong> sites. Hazard rating systems have<br />

been developed to grade f<strong>or</strong>est stands with respect to vulnerability to D. frontalis infestation (L<strong>or</strong>io 1980, Branham and<br />

Thatcher 1985, Mason et al. 1985). Several different methods have been devised. Each rating system has novel features, but<br />

all involve integration of a subset of variables: tree species, radial growth, height, and DBH; stand basal area (pine, hardwood,<br />

total), species composition, site index, and degree of crown closure; and landf<strong>or</strong>m classification. When applied to a<br />

f<strong>or</strong>est landscape, the hazard rating systems provide a general view of the distribution and abundance of host defenses against<br />

the insect. It is notew<strong>or</strong>thy that the hazard rating systems integrate a substantial knowledge base on the interaction of D.<br />

frontalis, host plants, and site conditions.<br />

To summarize, we have described the nature of the interaction of D. frontalis with host defenses and have identified<br />

several sources of variation imp<strong>or</strong>tant to population dynamics of the insect. Although the hazard rating systems have proven<br />

useful in defining vulnerability of f<strong>or</strong>est landscapes to herbiv<strong>or</strong>y by D. frontalis, they ign<strong>or</strong>e the seasonal dynamics of host<br />

defenses associated with different tree species, with individuals of the species, and with lightning-struck hosts. Below we<br />

will consider how the spatial and temp<strong>or</strong>al sources of variation in host defense can be integrated in the context of the natural<br />

hist<strong>or</strong>y of the insect. In effect, we view host defense capacity as a variable in landscape structure that can be characterized<br />

and defined. The distribution of <strong>this</strong> variable across the f<strong>or</strong>est landscape influences epidemiology of D. frontalis.<br />

Outbreaks and F<strong>or</strong>est Landscapes<br />

Herbiv<strong>or</strong>y at the landscape scale is significant because it is at <strong>this</strong> level of integration where the ecological and f<strong>or</strong>est<br />

management consequences of insect outbreaks are interpreted (Pickett and White 1985, F<strong>or</strong>man and Godron 1986, Barbosa<br />

and Schultz 1987, Turner 1987, Platt and Strong 1989, Turner and Gardner 1991, Holling 1992a, Wiens 1992, Coulson et al.<br />

1993). Advances in the geographic inf<strong>or</strong>mation system (GIS) technologies and the development of statistics f<strong>or</strong> analysis and<br />

description of spatially referenced data have greatly expedited studies of insect outbreaks in f<strong>or</strong>est landscapes (Liebold and<br />

Barrett 1993).<br />

Hist<strong>or</strong>ically, interest in epidemiology of D. frontalis has been primarily associated with the economic impact of the<br />

insect on f<strong>or</strong>est resources. Emphasis has focused on investigations of herbiv<strong>or</strong>y in the context of impact on f<strong>or</strong>est stand<br />

structure, i.e., the research has considered the effect of process (herbiv<strong>or</strong>y) on pattern (the configuration of elements that<br />

constitute the landscape mosaic). F<strong>or</strong> example, Fitzgerald et al. (1994) examined the relation of various suppression tactics,<br />

employed to modify D. frontalis populations, on the subsequent development of new infestations within the surrounding<br />

f<strong>or</strong>est landscape. 273


However, in pre-colonization f<strong>or</strong>ests of the southern US, herbiv<strong>or</strong>y by D. frontalis (along with fire) was likely<br />

involved in structuring the landscape (Schowalter et al. 198 l, Rykiel et al. 1988). As D. fr<strong>or</strong>_talis infestations are r_<strong>or</strong>mally<br />

suppressed in an eff<strong>or</strong>t to reduce economic loss by the insect, it has not been possible to evaluate <strong>this</strong> hypothesis un,it<br />

recently. Outbreaks of D. frontalis in Rare II Wilderness Areas on National F<strong>or</strong>ests in Texas have been allowed to follow a<br />

natural course, with only modest intervention to protect endangered species. The outbreaks, which occurred over a period of<br />

1983 to t994, have radically changed the pattern of the f<strong>or</strong>est landscape mosaic of these wilderness areas (Fig. t). Under <strong>this</strong><br />

epizootic condition, individual infestations coalesced and the pattern and extent of tree m<strong>or</strong>tality substantially modified the<br />

structure of the f<strong>or</strong>est landscape.<br />

Figure I .---The impact of herbiv<strong>or</strong>y by D. frontalis in pine f<strong>or</strong>es_ landscapes in east Texas. Under <strong>this</strong> epizootic condition,<br />

individual inl_stations coalesced and the pattern and extent of tree m<strong>or</strong>tality substantially modified the structure of<br />

the tbrest landscape.<br />

The alternate way of viewing insect outbreaks is to consider how landscape pattern affects the process of herbiv<strong>or</strong>y.<br />

Again, because D. frontalis infestations are n<strong>or</strong>mally suppressed, it was not possible in the past to examine <strong>this</strong> relation.<br />

Figure I clearly illustrates how the process of herbiv<strong>or</strong>y is influenced by the pattern of patches that were created by previous<br />

in|izstations of D. j}ontalis. There are few examples of empirical studies of the eflEcts of landscape pattern on the process of<br />

herbiv<strong>or</strong>y (Turner 1989).<br />

By definition, epidemiological studies are dynamic and theref<strong>or</strong>e the process of herbiv<strong>or</strong>y and the pattern of the tk_rest<br />

landscape mosaic change through mutual interaction, i.e., the pattern influences the process and the process influences the<br />

pattern. Although not investigated in a rig<strong>or</strong>ous manner, the scenario observed on the National F<strong>or</strong>est wilderness areas of<br />

Texas sugges{s that bark beetle herbiv<strong>or</strong>y drives the process of ecosystem succession (from the release to re<strong>or</strong>ganization<br />

phase in Holling's scheme 1992b) (see also Brown 1991).<br />

274


Functional Heterogeneity of F<strong>or</strong>est Landscapes: How Bark Beetles Perceive and Respond to Their Environment<br />

Studies of epidemiology begin with a consideration of issues associated with landscape structure. The term landscape<br />

structure refers to the spatial relationships between distinctive ecosystems (ecotopes, see Coulson et al. 1993), i.e., the<br />

distribution of"energy, materials, and species in relation to the sizes, shapes, numbers, kinds, and configurations of components<br />

(Turner 1989). Characteristics of landscape structure (p<strong>or</strong>osity of the f<strong>or</strong>est matrix, boundary configurations, patch size<br />

and shape, etc.) influence epidemiology of bark beetles. Structural characteristics of landscapes are often summarized and<br />

represented as heterogeneity. Although defined in several ways, the term heterogeneity is generally taken to mean composition<br />

of'parts of"different kinds. In considering the interaction of D. frontalis with the f<strong>or</strong>est landscape it is necessary to<br />

distinguish between measured and functional heterogeneity. Measured heterogeneity deals with physical features <strong>or</strong> components<br />

of a landscape. Functional heterogeneity deals with how an <strong>or</strong>ganism perceives and responds to its environment<br />

(Kolasa and Rollo 1991). There is obviously a direct link between patches, boundaries, and heterogeneity as boundaries<br />

define patches, and patchiness is what produces heterogeneity (Wiens 1992, Hansen and di Castri 1992).<br />

Tied to concepts associated with landscape heterogeneity is the subject of habitat structure. In general, habitat<br />

structure refers to the physical arrangement of objects in space (Bell et al. 1991). In studies of epidemiology, an essential<br />

task is to define the distribution, abundance, and location of habitat units. In effect, habitat structure is an element of measured<br />

heterogeneity, i.e., habitat can be viewed as a mappable element of a landscape mosaic. Southwood (1977) suggested<br />

that habitat serves as the template f<strong>or</strong> ecological strategies.<br />

Examination of how bark beetles perceive and respond to their environment at the landscape scale, i.e., evaluating<br />

functional heterogeneity, was difficult bef<strong>or</strong>e the availability of (i) GIS technologies, (it) spatially referenced databases, and<br />

(iii) spatial statistics. Also, until the text by F<strong>or</strong>man and Godron (1986), there was no standardized nomenclature f<strong>or</strong> describing<br />

landscapes from an ecological perspective (McIntosh 1991). With these problems partially solved, it is now possible to<br />

consider the substantial knowledge base on natural hist<strong>or</strong>y of bark beetles (and other <strong>or</strong>ganisms) in the context of the f<strong>or</strong>est<br />

landscapes where they occur (Dunning et al. 1992).<br />

In reference to D. frontalis, we have indicated that a great deal is known about host selection relative to the initiation<br />

and subsequent growth of infestations. In the case of initiation of infestations, lightning-struck hosts play a prominent role.<br />

F<strong>or</strong> infestation growth, variables associated with f<strong>or</strong>est stand hazard become significant. With knowledge of how these<br />

variables are spatially and temp<strong>or</strong>ally distributed, we can consider scenarios, based on the natural hist<strong>or</strong>y of the insect, that<br />

explain causes f<strong>or</strong> changes in the distribution and abundance of D. frontalis at the landscape scale.<br />

Ways and Means of Measuring Functional Heterogeneity<br />

Ways and means f<strong>or</strong> measuring landscape heterogeneity have been reviewed by Kolasa and Pickett (I991), Turner<br />

(1987, 1989), S. Turner et al. (1991), and M. Turner et al. (1991). Although there are several methodologies available, no<br />

single index is suitable f<strong>or</strong> characterization of functional heterogeneity f<strong>or</strong> a given <strong>or</strong>ganism. Each <strong>or</strong>ganism perceives and<br />

responds to its environment in unique ways. The elements of landscape structure that are imp<strong>or</strong>tant in defining herbiv<strong>or</strong>y by<br />

bark beetles, f<strong>or</strong> example, may be substantially different from those needed to characterize the process f<strong>or</strong> deer populations.<br />

Figure 2 illustrates the relation of functional and measured heterogeneity. In Figure 2 functional heterogeneity is represented<br />

as a gradient ranging from homogeneity on one end to heterogeneity on the other. Interpretation of the significance of index<br />

values in the mid-range of the gradient is difficult. We emphasize, also, that intermediate heterogeneity conditions are most<br />

difficult to interpret from the behavi<strong>or</strong>al perspective of the <strong>or</strong>ganism, i.e., it is difficult to predict how an <strong>or</strong>ganism will react<br />

to conditions of intermediate heterogeneity. F<strong>or</strong> <strong>this</strong> reason measurement of functional heterogeneity f<strong>or</strong> a specific <strong>or</strong>ganism<br />

will likely require use of several indices, each of which is sensitive to different aspects of landscape structure, e.g., connectivity,<br />

p<strong>or</strong>osity, fragmentation, boundary configuration, etc. To address <strong>this</strong> issue, we developed (<strong>or</strong> adapted) three different<br />

procedures to characterize functional heterogeneity of f<strong>or</strong>est landscapes relative to herbiv<strong>or</strong>y by D. frontalis. These indices<br />

are described briefly below.<br />

Indices of Heterogeneity<br />

Our initial goal in defining functional heterogeneity f<strong>or</strong> D. frontalis was to combine knowledge of natural hist<strong>or</strong>y of<br />

the insect with inf<strong>or</strong>mation on landscape structure. We developed (<strong>or</strong> modified) three indices to characterize functional<br />

heterogeneity. All use a weighted pattern-recognition computational approach, and each is sensitive to a different aspect of<br />

275


FUNCTIONAL VS. MEASURED HETEROGENEITY<br />

FunctionalHeterogeneity<br />

Measured Heterogeneity<br />

Figure 2.--Relationship of measured heterogeneity (composition of parts of different kinds) to functional heterogeneity (how<br />

an <strong>or</strong>ganism perceives and reacts to its environment). Note that measured heterogeneity f<strong>or</strong> the left and right matrices<br />

is the same, but the functional heterogeneity is different.<br />

Interpretation of the significance of index values in the mid-range of the gradient is difficult. The intermediate<br />

heterogeneity conditions are also the most difficult to interpret from the behavi<strong>or</strong>al perspective of the <strong>or</strong>ganism, i.e.,<br />

it is difficult to predict how an <strong>or</strong>ganism will react to conditions of intermediate heterogeneity.<br />

landscape pattern. Application of the procedures involves (i) classification of land elements based on ecological function, (ii)<br />

tiling of the landscape data (usually maps <strong>or</strong> images), and (iii) quantification of spatial pattern. Following is a brief description<br />

of each index along with an explanation of the calculation procedure. These methods apply f<strong>or</strong> any NxN array of<br />

weighted landscape elements. We are currently investigating the sensitivity of these indices.<br />

The Angular Moment of Inertia Index (AMI). The AMI is an index that is sensitive to variation in the dispersion<br />

of landscape elements. It is the n<strong>or</strong>malized product of a second moment calculation. The index is a measure of the dispersion<br />

of weighted matrix elements about the centroid of the landscape area. The AMI is calculated using the following<br />

procedures. First, the centroid of the weighted surface (matrix) is calculated. Because each element, a_j,within the matrix has<br />

a value and locational indices attached, the h<strong>or</strong>izontal (Xo)and the vertical (Yo) centroids can be calculated by<br />

Xo = (_j_i(aij),j) / _i_j aij<br />

Yo= (£_Zj(aij).i) / £,Ej aij<br />

Second, Xoand Yo are used to derive the second moment (M) about the centroid<br />

M = Z_jc(o. Xo)2 + (i - Yo)2) • au)<br />

However, because each element in the matrix represents a cell that has a real extent, the second moment f<strong>or</strong> each cell :<br />

must be calculated. The equation becomes<br />

276<br />

M = Z_i


The Connectivity Index (CI). The CI index is based on a run-length encoding strategy that is commonly used to<br />

encode raster (matrix) data. Run-length encoding is a good compression technique f<strong>or</strong> rasters that have a high degree of<br />

spatial autoc<strong>or</strong>relation. Run lengths are measured geometrically rather than arithmetically, giving m<strong>or</strong>e than linear value to<br />

longer runs. The computational procedures f<strong>or</strong> the CI index are given below.<br />

First, count the runs of matrix elements with the same value (i). F<strong>or</strong> example, 1l 11 would be a run of length L(1) =<br />

4. A row, column, <strong>or</strong> diagonal may have several runs. F<strong>or</strong> example, vect<strong>or</strong> 111222233333 has three runs. There can be<br />

multiple runs f<strong>or</strong> a given i.<br />

F<strong>or</strong> each row, column, <strong>or</strong> diagonal, we sum the runs f<strong>or</strong> each distinct element value. F<strong>or</strong> example, vect<strong>or</strong><br />

1112222211111 has three runs and two values. Converted to run values, the vect<strong>or</strong> is 123, 12345, 12345. The sum of runs,<br />

by element value, f<strong>or</strong> <strong>this</strong> vect<strong>or</strong> is ES(1) = 21 and ZS(2) = 15. F<strong>or</strong> a given run i of length L, the run value is S(i) = L * ( L +<br />

1 ) / 2. F<strong>or</strong> a given element value i, all possible S(i) are calculated by row, column, and diagonal, with the diagonal S(i)<br />

divided by 2.2 to compensate f<strong>or</strong> separation distance. The total run value f<strong>or</strong> a matrix and a given i is simply the ZS(i) f<strong>or</strong><br />

that matrix.<br />

F<strong>or</strong> the general NXN matrix, the smallest possible ZS(i), i = 1 to n where n is the number of different values, is<br />

ZSmin(i) = 2 ( N2+ N2/2 -2).<br />

ZSmin(i) exists where all run lengths are 1. The largest possible value, ZSmax(i), must be calculated using NXN<br />

matrix of constant i.<br />

Thus, the range is<br />

The unweighted CI is calculated as<br />

ESmax(i) = NZ(N+I) + 2(Z K(K+I))/2 z - N(N+I)/2 2<br />

R = ZSmax(i) - ZSmin(i).<br />

H = (ES(i) - ESmin(i)) / R.<br />

The CI f<strong>or</strong> a matrix where i values are relative to a function (i.e., hazard rating <strong>or</strong> probable use) can also be computed.<br />

The run values are calculated as weighted sums (WCI).<br />

WCI = 1; i(_:S(i)), i = 1,2 .....<br />

The Spatial Interaction Index. The third index of heterogeneity is based upon the gravity model and is sensitive to<br />

fragmentation of the highest valued landscape elements. When applied to geographic analysis, the gravity model is called<br />

spatial interaction. Theref<strong>or</strong>e, we have called <strong>this</strong> index spatial interaction (SI).<br />

SI is based on Newton's law of attraction between masses. It states that attraction is prop<strong>or</strong>tional to the product of<br />

two masses (assigned weight) and inversely prop<strong>or</strong>tional to the distance between them. F<strong>or</strong> a matrix of landscape elements<br />

represented in matrix f<strong>or</strong>m:<br />

, k = 1, where<br />

mkand ml are distinct elements of the matrix representing the attraction measure, r is the distance separating mk and m,, and t is<br />

the maximum value in the matrix.<br />

Functional Heterogeneity of F<strong>or</strong>est Landscapes and Host Defenses<br />

In <strong>this</strong> section we examine functional heterogeneity of f<strong>or</strong>est landscapes in relation to epidemiology of D. frontalis.<br />

Our approach is to provide (i) a general overview of the procedure used to characterize functional heterogeneity, (ii) illustrate<br />

how the methodology facilitates integration of intbrmation on landscape structure and natural hist<strong>or</strong>y of D. frontalis, and (iii)<br />

conclude with an interpretation of how host defenses influence epidemiology of the insect. 277


Approach f<strong>or</strong> Application of the Indices to Define Functional Heterogeneity of F<strong>or</strong>est Landscapes<br />

Use of the indices of heterogeneity is predicated on having a spatially referenced database that embraces the features<br />

of landscape structure essential to D. frontalis. Typically, development of GIS databases requires a substantial commitment<br />

to map and image processing in <strong>or</strong>der to identify and reference pertinent landscape variables. In the case ofD. frontalis, we<br />

are particularly interested in the habitat of the insect, i.e., patches of pines (graded by vulnerability) and lightning-struck<br />

hosts.<br />

In addition to the spatially referenced database, it is also necessary to define how the <strong>or</strong>ganism interacts with its<br />

environment, i.e., to relate the natural hist<strong>or</strong>y of the insect to habitat structure. We indicated earlier that a great deal is known<br />

about the interaction of D. frontalis and its hosts. Literature on <strong>this</strong> subject was reviewed in our treatme:nt of host defenses.<br />

An essential feature of D. frontalis/host interaction centers on the process of dispersal. The parameters f<strong>or</strong> <strong>this</strong> process<br />

determine how the netw<strong>or</strong>k of population centers (infestations) can be connected to habitat patches and lightning-struck<br />

hosts. Although the dispersal process as a whole is po<strong>or</strong>ly understood, certain elements of it have been studied in great detail,<br />

e.g., response of the insect to semiochemicals (Smith et al. 1993). We define boundaries f<strong>or</strong> the dispersal process, based on<br />

estimates available from the published literature (Coulson et al. 1979, Pope et al. 1980, Wagner et al. 1984).<br />

With the GIS database and inf<strong>or</strong>mation on how the insect interacts with its environment, we can use the indices to<br />

examine •functional heterogeneity of the f<strong>or</strong>est landscape. The general approach involves development of three types of<br />

maps: hazard, modified hazard, and heterogeneity. The hazard map is an abstract of the variables associated with different<br />

degrees of vulnerability of f<strong>or</strong>est stands to infestation by D. frontalis. The modified hazard map includes additional inf<strong>or</strong>mation<br />

on location of lighming-struck hosts, infestation centers, and insect behavi<strong>or</strong>. The heterogeneity map integrates these<br />

variables, i.e., inf<strong>or</strong>mation on habitat units, population centers, and insect behavi<strong>or</strong>. This map represents a view of how D.<br />

frontalis might interact with its environment. The map is produced by applying the indices of heterogeneity to the modified<br />

hazard map. A moving-window calculation (Fig. 3) is used. The moving-window approach is a common technique in<br />

landscape ecological studies (Turner et al. 1991). It is particularly useful in defining functional heterogeneity, as window<br />

size and shape can be changed to accommodate different spatial scales (Weins 1989, Levin 1992) (Fig. 4). In the particular<br />

application, we are interested in the spatial scale used by D. frontalis, which is largely set by the distribution and abundance<br />

of habitat units and population centers and the dispersal capability of the insect.<br />

MOVING-WINDOW ANALYSIS<br />

J - . n<br />

i I J' +1 n-1<br />

-{ii{ i i<br />

: ,<br />

........ 1 ....... !...........<br />

---Inaut Array .... Output Array--<br />

.... _<br />

L, .....<br />

Figure 3.--The moving-window analysis applied to an array. The arrays can be raster renditions of landscape maps. Each cell<br />

(i,j) represents a 2-d sample of the landscape. In <strong>this</strong> example the window is 3X3 cells. In our methodology functional<br />

heterogeneity is calculated f<strong>or</strong> land attributes in the windows of the input array and placed in the output array.<br />

278<br />

n-1


FUNCTIONAL HETEROGENEITYANALYSIS :<br />

A TECHNIQUE THAT USES BOTH PATTERNAND SCALE<br />

MEASURED LANDSCAFE FUNCTIONAL LANDSCAPES<br />

Moving-Window Application of<br />

the Weighted Connectivity Index<br />

1 2<br />

* grayshadesindicatemapped • sizeof movingwindowc<strong>or</strong>respondsto scaleof energy,<br />

landscapezones material,<strong>or</strong><strong>or</strong>ganismflows<br />

, grayshadesindicateindexvalues<br />

.... ;IFT]<br />

1, closelyresemblesmeasured :3 ° vaguelyresemblesmeasured<br />

landscape landscape<br />

°onlyboundariesare ° indexvaluesspatiallyvaried<br />

heterogeneous ° landscapeis functionally<br />

heterogeneous<br />

Figure 4.--Functional heterogeneity analysis using pattern recognition alg<strong>or</strong>ithms and moving-window calculations.<br />

Relating Functional Heterogeneity to Epidemiology of D. frontalis<br />

Although D. frontalis is among the most th<strong>or</strong>oughly studied pest species, there is no mechanistic explanation f<strong>or</strong> how<br />

the natural hist<strong>or</strong>y of the insect operates at the landscape scale to produce observed patterns of epidemiology (Dunning et al.<br />

1992). We suggest that bark beetle epidemiology in a pine f<strong>or</strong>est landscape involves a netw<strong>or</strong>k of [i] lightning-struck hosts<br />

(Lovelady 1994), which serve as sinks during dispersal (Pullian 1988, Pullian and Danielson 1991); [ii] previous existing<br />

infestations, which serve both as sources and sinks; and [iii] high hazard stands, which also serve as sinks and are needed f<strong>or</strong><br />

development of infestations. With only modest requirements f<strong>or</strong> dispersal distance (1 kin), the bark beetles can reduce high<br />

measured heterogeneity in pine f<strong>or</strong>est landscapes into high functional homogeneity, i.e., the insects can link dispersed food<br />

and habitat resources. Specific attributes of landscape structure will influence how <strong>this</strong> scenario is played out in different<br />

f<strong>or</strong>est environ ments.<br />

In the following example, we examine how D. frontalis interacts within a meso-scale (100 to 1,000,000 ha) f<strong>or</strong>est<br />

landscape. The GIS-based study (McFadden 1994) was conducted on the Sam Houston National F<strong>or</strong>est in southeast Texas.<br />

This landscape is vegetated primarily with loblolly, Pinus taeda L., and sh<strong>or</strong>tleaf, P echinata Mill, pines and mixed hardwood<br />

species. Populations of D. frontalis in the f<strong>or</strong>est have cycled from enzootic to epizootic levels f<strong>or</strong> m<strong>or</strong>e than two<br />

decades.<br />

279


Data f<strong>or</strong> the study were acquired from thematic maps, aerial photographs, lightning-strike rec<strong>or</strong>ds, and f<strong>or</strong>est management<br />

rec<strong>or</strong>ds. The thematic map inf<strong>or</strong>mation included: ownership boundaries, roads, utility lines, hydrology, and f<strong>or</strong>est<br />

compartment boundaries. Locations of D. frontalis infestations were obtained by ge<strong>or</strong>eferencing and interpreting aerial<br />

photographs, using a computer-assisted mapping system called MIPS (Map and Image Processing System). Aerial photographs<br />

were provided by the <strong>USDA</strong> F<strong>or</strong>est Service. Once located on aerial photographs, the UTM co<strong>or</strong>dinates f<strong>or</strong> infestation<br />

centers were collected and entered into the GIS (ARC/INFO) as point locations. Lightning strike data f<strong>or</strong> the region were<br />

purchased fi'om R-Scan C<strong>or</strong>p<strong>or</strong>ation. Again, UTM co<strong>or</strong>dinates f<strong>or</strong> lightning strikes were entered as point locations. Inf<strong>or</strong>mation<br />

on tk_reststands was taken from a database (CISC-Continuous Invent<strong>or</strong>y of Stand Conditions), which is maintained by<br />

the F<strong>or</strong>est Service. These data sources were used to develop a sequence of three maps: the hazard, the modified hazard, and<br />

the heterogeneity. Each map is described below.<br />

The Hazard Map. First in the sequence is a general D. frontalis hazard map (Fig. 5). This map was developed by<br />

hazard rating l<strong>or</strong>est stands in the study area. In <strong>this</strong> specific application, we used tree species and age as the principal<br />

variables of hazard. We mentioned previously that several hazard rating systems have been developed (L<strong>or</strong>io 1980). The<br />

utility of <strong>this</strong> map is that it defines patches in the f<strong>or</strong>est matrix that are suitable f<strong>or</strong> the development of infestations, given the<br />

presence of the insect (and acceptable weather conditions). The map also reflects habitat patches graded by host defense<br />

capability, i.e., the high hazard stands have the lowest capacity f<strong>or</strong> host defense. This rendition of the landscape will have<br />

seasonal trend reflecting variation in host defenses associated with the n<strong>or</strong>mal physiological changes associated with tree<br />

phenological development described by L<strong>or</strong>io (1988). The general hazard map will change slowly, i.e., it will require<br />

periodic updating, perhaps every 2 <strong>or</strong> 3 years, to remain an accurate p<strong>or</strong>trayal of hazard.<br />

5 ecies Ty e Hazard Rating<br />

m_h<strong>or</strong>tleaf _White [] No Hazarct<br />

Pine /HiGk<strong>or</strong>y<br />

Loblolly Pine/ [] Low Hazar_i<br />

ne mHardwo, [] Medium Hazar_<br />

Mix<br />

DSweetguml [] High Hazar_<br />

Willow BWater<br />

Figure 5._Maps of a :f<strong>or</strong>estlandscape classified by species types and hazard to herbiv<strong>or</strong>y by D. fi'ontalis. Several systems<br />

f<strong>or</strong> hazard rating have been developed. All involve integration of a subset of variables: tree species, radial growth,<br />

height, and DBH; stand basal area (pine, hardwood, total), species composition, site index, and degree of crown<br />

closure; and landf<strong>or</strong>m classification. When applied to a f<strong>or</strong>est landscape, the hazard rating systems provide a general<br />

view of the distribution and abundance of host defenses against the insect.<br />

280


The Modified Hazard Map. The next level of complexity in the map sequence involves adding inf<strong>or</strong>mation on the<br />

location of lightning-struck hosts and existing infestations (Fig. 6). The lightning-struck hosts represent sinks f<strong>or</strong> dispersing<br />

beetles. We indicated above that these hosts have greatly reduced defense capacity, and they can be located, presumably<br />

through olfact<strong>or</strong>y cues, by D. frontalis. The role of lightning-struck hosts has been examined in detail in Coulson et al.<br />

(1986a), Flamm et al. (1992), Lovelady et al. (1991), and Lovelady (1994). The existing infestations can serve both as sinks<br />

and sources f<strong>or</strong> dispersing beetles. The product of adding the lightning-struck hosts and infestation centers to the hazard is<br />

illustrated in Figure 7 (left panel). The modified hazard map also inc<strong>or</strong>p<strong>or</strong>ates inf<strong>or</strong>mation on dispersal distance of the insect.<br />

We assumed that the insect could disperse a distance of ca 1 krn. This conservative estimate was based on rep<strong>or</strong>ts from the<br />

literature cited above. On the modified hazard map (Fig. 7 - left panel), the dispersal distance was represented as an area (a<br />

circle with a 1 kin radius) around the lightning-struck hosts and infestation centers. The lightning-struck hosts are point<br />

locations in the f<strong>or</strong>est matrix. However, these hosts have a substantial impact on epidemiology when their locations are<br />

considered within the context of the dispersal range ofD. frontalis. The modified hazard map contains the elements of<br />

landscape structure involved in epidemiology, i.e., hazard rated f<strong>or</strong>est stands, lightning centers, and population centers. The<br />

addition of dispersal distance provides the means f<strong>or</strong> connecting the elements. This rendition of the landscape has a significant<br />

weather-related annual pattern. Thunderst<strong>or</strong>m (and thus lightning) activity follows seasonal cycles (Lovelady 1994).<br />

Insect development and dispersal behavi<strong>or</strong> also have seasonal trends. The inf<strong>or</strong>mation content of the modified hazard map<br />

will change during the course of a year, i.e., it will require frequent updating, perhaps weekly <strong>or</strong> monthly, to remain an<br />

accurate rendition of hazard,<br />

Lightning Infestations<br />

June 1989 June 1989<br />

Lightning & SPB<br />

A Strikes Infestations<br />

S_and<br />

Stan_ _ 13oundarie5<br />

A Boundaries<br />

Figure 6. Maps of a f<strong>or</strong>est landscape with point locations f<strong>or</strong> lightning-struck hosts and existing infestation centers. Lightning-struck<br />

hosts have greatly reduced defense capacity and are used as refuges by D. frontalis, i.e., they are sinks f<strong>or</strong><br />

dispersing populations of the insect. The infestation centers serve as both sinks and sources f<strong>or</strong> dispersing populations<br />

of the insect.<br />

281


Modified Stand Hazard Ma Here ene t<br />

Figure 7. Modified hazard map (left panel) inc<strong>or</strong>p<strong>or</strong>ates in:f<strong>or</strong>mation on lightning-struck hosts and existing infestation<br />

centers. This map also contains inf<strong>or</strong>mation on the dispersal behavi<strong>or</strong> of D. ¢?o_talis. The dispersal distance was<br />

represented as an area (a circle with a 1 km radius) around the lightning-struck hosts and infestation centers. Dis-<br />

persal and host selection behavi<strong>or</strong> of the insect link the netw<strong>or</strong>k of high hazard stands, infestation centers, and<br />

lightning-struck host. The heterogeneity map (right panel) was developed using the AMI index applied to the<br />

modified hazard map, using the moving-window calculation. This map illustrates how the netw<strong>or</strong>k of high hazard<br />

:stands, liglhtning-struck hosts, and existing population centers is connected through dispersal behavi<strong>or</strong> of the insect.<br />

The Heterogeneity Map. Third in the sequence is the functional heterogeneity map (Fig. 7 - right panel). In <strong>this</strong><br />

particular example, tile AMI index was applied to the modified hazard map, using the moving-window calculation described<br />

above. Recall that the AMI index is sensitive to dispersion of landscape elements. The functional heterogeneity index allows<br />

f<strong>or</strong> integration of the inf<strong>or</strong>mation on landscape structure and insect behavi<strong>or</strong> known to influence epidemiology of D. frontatis.<br />

The heterogeneity map (Fig. 7-right panel) is a visualization of the results. This map illustrates how the netw<strong>or</strong>k of high<br />

hazard stands_ lightni_g.-struck hosts, and existing population centers is connected through dispersal behavi<strong>or</strong> of the insect.<br />

Knowledge of <strong>this</strong> interaction will certainly increase with new research on epidemiotogy, and <strong>this</strong> added complexity can be<br />

accommodated by the approach, Figure 7 (right panel) is a characterization of only one point in time. During the course of<br />

the year, functional heterogeneity of the landscape will change f<strong>or</strong> the reasons outlined above. Functional heterogeneity will<br />

also change during successive years of an insect outbreak, as herbiv<strong>or</strong>y creates additional fragmentation of the f<strong>or</strong>est matrix.<br />

Furtherm<strong>or</strong>e, there will be a flux in habitat available to the insect in successive years, as high hazard stands are depleted<br />

through herbiv<strong>or</strong>y and new stands grow into the vulnerable age classes. The index values are a quantitative measure of<br />

ftmctional heterogeneity and theref<strong>or</strong>e can be used f<strong>or</strong> tile analysis and description of sequential map data. Recall, also, that<br />

the tt_ree indices are sensitive to dilYerent aspects of landscaPe pattern. Therefk)re, each index will provide alternative insights<br />

into how O_ ._f?ontatis perceives and reacts to its environment.<br />

Host Defenses and Epidemiology of Dendroctonus frontalis<br />

"Ik_summarize, at the onset of <strong>this</strong> study, we indicated that epidemiology of D. frontalis was related to the distribution<br />

of host defenses in f<strong>or</strong>est landscapes. Previous research had clearly established that infestations typically occur in high<br />

hazard stands consisting of hosts with reduced capacity f<strong>or</strong> defense against colonization by bark beetles. Hazard rating<br />

282


systems facilitate mapping of f<strong>or</strong>ests acc<strong>or</strong>ding to their vulnerability to bark beetle infestation. These hazard maps define a<br />

landscape mosaic comprised of f<strong>or</strong>est stands graded by defense capacity (Fig. 5). Typically, the high hazard stands are also<br />

the preferred habitat of the insect, i.e., populations of bark beetles are clustered in high hazard stands. As long as population<br />

density remains sufficient to overcome host defenses, infestations grow in size through colonization of neighb<strong>or</strong>ing hosts.<br />

When population density falls below the level necessary f<strong>or</strong> successful host colonization, the insects must disperse to find<br />

suitable refuges. Two strategies are likely. The insects can disperse to supplement populations in existing infestations at<br />

other locations <strong>or</strong> they can colonize lightning-struck hosts, which have greatly reduced defense capacity. These two elements,<br />

existing infestations and lightning-struck hosts, add to the picture of host defense f<strong>or</strong> the f<strong>or</strong>est landscape (Fig. 7-1eft<br />

panel). However, f<strong>or</strong>est landscapes in the southeastern US are often highly fragmented, and the vulnerable stands, existing<br />

infestations, and lightning-struck hosts are dispersed. Behavi<strong>or</strong>al strategies associated with dispersal and host selection link<br />

populations of the insect to suitable habitat centers (Fig. 7-1eft panel) (Southwood 11977). Finally, the indices of functional<br />

heterogeneity serve to integrate the inf<strong>or</strong>mation on the distribution of host defenses with knowledge on how the insect<br />

perceives and responds to its environment (Fig. 7-right panel). Further insight into mechanisms of epidemiology of D.<br />

frontalis can be gained through analysis of changes in functional heterogeneity of sequential landscape scenes.<br />

SUMMARY<br />

This paper focused on functional heterogeneity of f<strong>or</strong>est landscapes and epidemiology of D. frontalis. Specific conclusions<br />

from the study include the following:<br />

1. The general premise of <strong>this</strong> study was that epidemiology of D. frontalis involves a netw<strong>or</strong>k of high hazard stands,<br />

lightning-struck hosts, and existing population centers. This netw<strong>or</strong>k is connected at the landscape scale through dispersal<br />

behavi<strong>or</strong> of the insect.<br />

2. The concept of functional heterogeneity (how an <strong>or</strong>ganism perceives and responds to its environment) provides a<br />

useful way to <strong>or</strong>ganize knowledge about landscape structure and insect behavi<strong>or</strong>. Although we have focused on functional<br />

heterogeneity of f<strong>or</strong>est landscapes and epidemiology of D. frontalis, the approach can be used f<strong>or</strong> other insects and taxa.<br />

3. The :indices of functional heterogeneity were found to be useful in integrating the landscape variables known to<br />

influence epidem iology of D. frontalis.<br />

4. Measurements of functional heterogeneity of f<strong>or</strong>est landscapes inc<strong>or</strong>p<strong>or</strong>ate two types of knowledge about host defenses,<br />

The first deals with knowledge embedded in the various types of hazard rating systems used to grade f<strong>or</strong>est stands by<br />

vulnerability to D. J?oHtalis. The second deals with knowledge on the interaction of lighting-struck hosts and D. frontalis. The<br />

functional heterogeneity map provides a visualization of the distribution of host defenses within a meso-scale f<strong>or</strong>est landscape.<br />

5. The functional heterogeneity indices are a tool that can be used to investigate several fundamental issues in epidemiology.<br />

F<strong>or</strong> example we hypothesize that (i) functional heterogeneity of f<strong>or</strong>est landscapes influences the rate and extent of<br />

herbiv<strong>or</strong>y by bark beetles and that (ii) initiation, growth, and decline phases of f<strong>or</strong>est insect outbreaks can be predicted by<br />

definition of the change points in functional heterogeneity of landscapes.<br />

LITERATURE CITED<br />

BARBOSA, R and SCHULTZ, J.C., eds. 1987. Insect Outbreaks. Academic Press, San Diego, CA.<br />

BRANHAM, S J., and THATCHER, R.C., eds. 1985. Proc. Integrated Pest Management Symposium. Gen. Tech. Rep. SO-<br />

56. New Orleans, LA: U.S. Department of Agriculture, F<strong>or</strong>est Service.<br />

BELl.,, S,S._ McCOY, ED., and RUSHINSKY, H.R., eds. 1991. Habitat Structure. Chapman and Hall, NY.<br />

BROWN, V.K. 1991. The effects of changes in habit structure during seccessin in terresterial communities. In Bell, S.S.,<br />

McCoy, E.D., and Rushinsky, H.R., eds. 1991. Habitat Structure. Chapman and Hall, NY.<br />

283


COULSON, R.N., PULLEY, RE., POPE, D.N., FARGO, W.S., GAGNE, J.A., and KELLEY, C.L. 1979. Estimation of<br />

survival and allocation of adult Dendroctonusfrontalis between trees during the development of an infestation. In<br />

Berryman, A.A. and Safranyik, L., eds. Dispersal of F<strong>or</strong>est Insects: Evaluation, The<strong>or</strong>y, and Management Implications.<br />

Proc. IUFRO Syrup., Sandpoint, ID.<br />

COULSON, R.N., HENNIER, RB., FLAMM, R.O., RYKIEL, E.J., HU, L.C., and PAYNE, T.L. 1983. The role of lightning<br />

in the epklemiology of the southern pine beetle. Ziet. Ang. Entomol. 96: 182-93.<br />

COULSON, R.N., RYKIEL, E.J., SAUNDERS, M.C., PAYNE, T.L., FLAMM, R.O., WAGNER, T.L., and HENNIER, RB.<br />

1985a. A conceptual model of the role of lightning in the epidemiology of the southern pine beetle. In Safranyik, L.,<br />

ed. The Role of the Host in Population Dynamics of F<strong>or</strong>est Insects. Proc. IUFRO Syrup., Banff, Canada.<br />

COULSON, R.N., FLAMM, R.O., WAGNER, T.L., RYKIEL, EJ., SHARPE, P.J.H., PAYNE, T.L., and LIN, S.K. t985b.<br />

Population dynamics of initiation and growth of southern pine beetle infestations. In Branham, S.J. and Thatcher,<br />

R.C., eds. Proc. Integrated Pest Management Symposium. Gen. Tech. Rep. SO-56. New Orleans, LA: U.S. Department<br />

of Agriculture, F<strong>or</strong>est Service.<br />

COUt,SON, R.N., FLAMM, R.O., PULLEY, RE., PAYNE, T.L., RYKIEL, E.J., and WAGNER, T.L. 1986a. Response of<br />

the southern pine bark beetle guild to host disturbance. Environ. Entomol. 15:859-868.<br />

COULSON, R.N., RYKIEL, E.J., and CROSSLEY, D.A., Jr. 1986b. Activities of insects in f<strong>or</strong>ests: implications f<strong>or</strong> wilderness<br />

area management. In Kulhavy, D.L. and Conner, R.N., eds. Wilderness and Natural Areas in the East: A Management<br />

Challenge. Stephen E Austin State Univ., Nacogdoches, TX.<br />

COULSON, R.N., FITZGERALD, J.W., SAUNDERS, M.C., and OLIVERIA, F.L. 1993. Spatial analysis and integrated<br />

pest management in a landscape ecological context. In Liebold, A.M. and Barrett, H.R., eds. Proceedings: Spatial<br />

Analysis and F<strong>or</strong>est Pest management. Gen. Tech. Rep. NE-175. Radn<strong>or</strong>, PA: U.S. Department of Agriculture,<br />

F<strong>or</strong>est Service.<br />

DUNNING, J.B., DANIELSON, B.J., and PULLIAM, H.R. 1992. Ecological processes that affect populations in complex<br />

landscapes. Oikos 65. 169-75.<br />

FITZGERALD, J.W., COULSON, R.N., PULLEY, RE., FLAMM, R.O., OLIVERIA, F.L., SWAIN, K.M., and<br />

DRUMMOND, D.B. 1994. Suppression tactics _brDendroctonusfrontalis Zimmermann (Coleoptera: Scolytidae):<br />

an examination of the development of infestations adjacent to treatment sites. J. Econ. Entomol. (In press).<br />

FLAMM, R.O. and COULSON, R.N. 1988. Population dynamics of the southern pine bark beetle guild in traumatized hosts.<br />

In Mattson, W. J., Levieux, J., and Bernard-Dagan, C., eds. Plant Resistance Mechanisms to Insects and Pathogens.<br />

Spri nger-Verlag, NY.<br />

FLAMM, R.O., COULSON, R.N., and PAYNE, T.L. 1988. The southern pine beetle. In Berryman, A.A., ed. Dynamics of<br />

F<strong>or</strong>est Insect Populations. Plenum Press, NY.<br />

FLAMM, R.O., PUI,LEY, RE., and COULSON, R.N. 1992. Colonization of disturbed trees by the southern pine bark beetle<br />

guild (Coleoptera: Scolytidae). Environ. Entomol. 22: 62-70.<br />

FORMAN, R.T.T. and GODRON, M. 1986. Landscape Ecology. John Wiley and Sons, NY.<br />

HANSEN, A.J. and di CASTRI, F., eds. 1992. Landscape Boundaries: consequences f<strong>or</strong> biotic diversity and ecological<br />

flows. Ecol. Studies 92. Springer-Verlag, NY.<br />

HOLMNG, C.S. 1992a. The role of f<strong>or</strong>est insects in structuring the b<strong>or</strong>eal landscape. In Shugart, H.H., Leemans, R., and<br />

Bonan, G.B., eds. A Systems Analysis of the Global B<strong>or</strong>eal F<strong>or</strong>est. Cambridge Univ. Press, Cambridge.<br />

HOLLING, C. S. 1992b. Cross-scale m<strong>or</strong>phology, geometry, and dynamics of ecosystems. Ecol. Monographs 62: 447-502.<br />

284<br />

) ¸7¸¸7 :!;<br />

:


KOLASA, J. and ROLLO, C.D. 1991. Introduction: The heterogeneity of heterogeneity: a glossary. In Kolasa, J. and<br />

Pickett, S.T.A., eds. Ecological Heterogeneity. Ecological Studies 86. Springer-Verlag, NY.<br />

KOLASA, J. and PICKETT, S.T.A., eds. 1991. Ecological Heterogeneity. Ecol. Studies 86. Springer-Verlag, NY.<br />

LEVIN, S.A. 1992. The problem of pattern and scale in ecology. Ecology 73: 1943-67.<br />

LIEBOLD, A.M. and BARRETT, H.R., eds. 1993. Proceedings: Spatial Analysis and F<strong>or</strong>est Pest management. Gen. Tech.<br />

Rep. NE-175. Radn<strong>or</strong>, PA: U.S. Department of Agriculture, F<strong>or</strong>est Service.<br />

LORIO, RL., Jr. 1980. Rating stands f<strong>or</strong> susceptibility to SPB. In Thatcher, R C., Searcy, J.L., Coster, J.E., and Hertel,<br />

G.D., eds. The Southern Pine Beetle. Tech. Bull. 1631. Washington, DC: U.S. Department of Agriculture, F<strong>or</strong>est<br />

Service and Science and Education Administration.<br />

LORIO, RL., Jr. 1988. Growth and differentiation balance relationships in pines affect their resistance to bark beetles<br />

(Coleoptera: Scolytidae). In Mattson, W.J., Levieux, J., and Bernard-Dagan, C., eds. Mechanisms of Woody Plant<br />

Defenses Against Insects: Search f<strong>or</strong> Pattern. Springer Verlag, NY.<br />

LOVELADY, C.N., PULLEY, RE., COULSON, R,N., and FLAMM, R.O. 1991. Relation of lightning to herbiv<strong>or</strong>y by the<br />

southern pine bark beetle guild (Coleoptera: Scolytidae). Environ. Entomol. 20: 1279-1284.<br />

LOVELADY, C.N. 1994. Landscape Study of Bark Beetle Herbiv<strong>or</strong>y and the Lightning Disturbance Regime. Ph.D.<br />

Dissertation, Texas A&M University. 127 p.<br />

MASON, G.N., LORIO, RL., Jr., BELANGER, R.R, and NETTLETON, W.A. 1985. Rating the susceptibility of stands to<br />

southern pine beetle attack. Agric. Handb. 645. Washington, DC: U.S. Department of Agriculture, F<strong>or</strong>est Service.<br />

McFADDEN, B.A. 1994. Predicting F<strong>or</strong>est Insect Outbreaks: The Role of Lightning in the Epidemiology of the Southern<br />

Pine Beetle. MS Thesis, Texas A&M University (In prep.).<br />

McINTOSH, R.R 1991. Concept and terminology of homogeneity and heterogeneity in ecology, Chapter 2 (p. 23-46). In<br />

Kolasa, J. and Pickett, S.T.A., eds. Ecological Heterogeneity. Ecol. Studies 86. Springer-Verlag, NY.<br />

PICKETT, S.T.A. and R S. WHITE, RS., eds. 1985. The Ecology of Natural Disturbance and Patch Dynamics. Academic<br />

Press, NY. 472 p.<br />

PLATT, W.J. and STRONG, D.R., eds. 1989. Gaps in f<strong>or</strong>est ecology (special feature). Ecology 70: 535-76.<br />

POPE, D.N., COULSON, R.N., FARGO, W.S., GAGNE, J.A., and KELLEY, C.W. 1980. The allocation process and<br />

between-tree survival probabilities in Dendroctonusfrontalis infestations. Res. Pop. Ecol. 22: 197-210.<br />

PULLIAN, H.R. 1988. Sources, sinks, and population regulation. Am. Nat. 132: 652-661.<br />

PULLIAN, H.R and DANIELSON, B.J. 1991. Sources, sinks, and habitat selection: a landscape perspective on population<br />

dynamics. Am. Nat. 137: 550-566.<br />

RYKIEL, E.J., Jr., COULSON, R.N., SHARPE, RJ.H., ALLEN, T.F.H., and FLAMM, R.O. 1988. Disturbance propagation<br />

by bark beetles as an episodic landscape phenomenon. J. Land. Ecol. 1: 129-39.<br />

SCHOWALTER, T.D., COULSON, R.N., and CROSSLEY, D.A., Jr. 1981. Role of southern pine beetle and fire in maintenance<br />

of structure and function of the southeastern coniferous f<strong>or</strong>est. Environ. Entomol. 10:821-825.<br />

SMITH, M.T., SALOM, S.M., and PAYNE, T.L. 1993. The Southern Pine Bark Beetle Guild: An Hist<strong>or</strong>ical Review of<br />

<strong>Research</strong> on the Semiochemical-Based Communication System of the Five Principal Species. Virginia Agric.<br />

Experiment Stn. Bulletin 93-4.<br />

285


SOUTHWOOD, T.R.E. 1977. Habitat, the templet f<strong>or</strong> ecological strategies. J. Anita. Ecol. 46: 337-65.<br />

TURNER, M.G., ed. 1987. Landscape Heterogeneity and Disturbance. Ecol. Studies 65. Springer-Verlag, NY ....<br />

TURNER, M.G. 1989. Landscape ecology: the effect of pattern on process. Annu. Rev. Ecology and Systematics. 20: 11711-<br />

97.<br />

TURNER, M.G. and GARDNER, R.H., eds. 1991. Quantitative Methods in Landscape Ecology. Ecological Studies 82.<br />

Springer-Verlag, NY. t<br />

TURNER, S.J., O'NEIL, R.V., CONLEY, W., CONLEY, M.R., and GROVER, H.D. 1991. Pattern and scale: statistics f<strong>or</strong> l<br />

landscape ecology. In Turner, M.G. and Garner, R.H., eds. Quantitative Methods in Landscape Ecology. Ecological<br />

Studies 82. Springer-Verlag, NY. 536 p.<br />

WAGNER, T.L., GAGNE, J.A., SHARPE, EJ.H., and COULSON, R.N. 1984. Effects of constant temperature on longevity<br />

of adult southern pine beetles (Coleoptera: Scolytidae). Environ. Entomol. 13:1125-1130.<br />

WIENS, J.A. 1989. Spatial scaling in ecology. Functional Ecology 3: 385-97. _<br />

WIENS, J.A. 1992. Ecological flows across landscape boundaries: a conceptual overview, Chapter 10 (p. 217-35). In _<br />

Hansen, A.J. and di Castri, E, eds. Landscape Boundaries: consequences f<strong>or</strong> biotic diversity and ecological flows. _<br />

Ecological Studies 92. Springer-Verlag, NY.<br />

ACKNOWLEDGMENTS<br />

We acknowledge and thank Audrey M. Bunting f<strong>or</strong> assistance in the preparation of the figures and text of <strong>this</strong> paper.<br />

We are grateful to the <strong>USDA</strong> F<strong>or</strong>est Service, National F<strong>or</strong>ests in Texas f<strong>or</strong> providing access to GIS databases and f<strong>or</strong>est<br />

management rec<strong>or</strong>ds; and to the Texas F<strong>or</strong>est Service f<strong>or</strong> providing access to their operational rec<strong>or</strong>ds on D. frontalis. The<br />

w<strong>or</strong>k rep<strong>or</strong>ted herein was supp<strong>or</strong>ted by the <strong>USDA</strong> Competitive Grants Program [TEX06961 (Landscape Study of Bark Beetle<br />

Herbiv<strong>or</strong>y and the Lightning Disturbance Regime) and TEX08185 (Predicting F<strong>or</strong>est Insect Outbreaks)], and by the Texas<br />

Agricultural Experiment <strong>Station</strong>. This paper is TA No. 31482 of the Texas Agricultural Experiment <strong>Station</strong>.<br />

286<br />

_U.S. _ PRINTING OFFICE: 1996-757--939


) i,i_i)i_)i_i_<br />

U.S. Departme_]t of Agriculture, F<strong>or</strong>est Service.<br />

] 996. M at tsot_, \Vi ]1iat-1_J,; N ie :_ni1_i,P ekka; Ross i, M at ti, eds. Dy ha,hies o f f<strong>or</strong>es t<br />

i!f_sll herbi_,<strong>or</strong>y: Quest f<strong>or</strong> patter,, and principle. (3c,,. I>ch. Rep. NC-183. SI. Paul,<br />

MN: LLS, Department of AgriculRare, F<strong>or</strong>est Service, N<strong>or</strong>th Central F<strong>or</strong>csi Experi-<br />

meat <strong>Station</strong>. 286 p.<br />

Herbivo U on woody phmts is highly var'ktble in both space and time. This<br />

: proceedings addresses one of its root causes, the highly intricate and dynamic<br />

relationships that exist between most herbiv<strong>or</strong>es and their hosts plants. It emphasizes<br />

t:ha_ the conseque_-_ces of herbiv<strong>or</strong>y both to the consumer and to the producer plant<br />

oi_e_ balance on a raz<strong>or</strong>'s edge .......... depending o_?the exact timing of herbiv<strong>or</strong>e am_cks,<br />

... ;.l_,.t _he specific plant tissues being injured. Herbiv<strong>or</strong>y also varies substantially<br />

amo_'_g ii*_divMual pkmts in relationship to the inherent resistance/susceptibility of<br />

_/_ it_dividual plants ........ which itself heavily depends on the patticutar physical and biotic<br />

(c

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!