Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Background and aims

Distribution shifts and natural selection during past climatic changes are important factors in determining the genetic structure of forest species. In particular, climatic fluctuations during the Quaternary appear to have caused changes in the distribution ranges of plants, and thus strongly affected their genetic structure. This study was undertaken to identify the responses of the conifer Cryptomeria japonica, endemic to the Japanese Archipelago, to past climatic changes using a combination of phylogeography and species distribution modelling (SDM) methods. Specifically, this study focused on the locations of refugia during the last glacial maximum (LGM).

Methods

Genetic diversity and structure were examined using 20 microsatellite markers in 37 populations of C. japonica. The locations of glacial refugia were assessed using STRUCTURE analysis, and potential habitats under current and past climate conditions were predicted using SDM. The process of genetic divergence was also examined using the approximate Bayesian computation procedure (ABC) in DIY ABC to test the divergence time between the gene pools detected by the STRUCTURE analysis.

Key results

STRUCTURE analysis identified four gene pools: northern Tohoku district; from Chubu to Chugoku district; from Tohoku to Shikoku district on the Pacific Ocean side of the Archipelago; and Yakushima Island. DIY ABC analysis indicated that the four gene pools diverged at the same time before the LGM. SDM also indicated potential northern cryptic refugia.

Conclusions

The combined evidence from microsatellites and SDM clearly indicates that climatic changes have shaped the genetic structure of C. japonica. The gene pool detected in northern Tohoku district is likely to have been established by cryptic northern refugia on the coast of the Japan Sea to the west of the Archipelago. The gene pool in Yakushima Island can probably be explained simply by long-term isolation from the other gene pools since the LGM. These results are supported by those of SDM and the predicted divergence time determined using ABC analysis.

Free full text 


Logo of annbotAboutAuthor GuidelinesEditorial BoardAnnals of Botany
Ann Bot. 2014 Dec; 114(8): 1687–1700.
Published online 2014 Oct 29. https://doi.org/10.1093/aob/mcu197
PMCID: PMC4649686
PMID: 25355521

Evidence for cryptic northern refugia in the last glacial period in Cryptomeria japonica

Abstract

Background and Aims

Distribution shifts and natural selection during past climatic changes are important factors in determining the genetic structure of forest species. In particular, climatic fluctuations during the Quaternary appear to have caused changes in the distribution ranges of plants, and thus strongly affected their genetic structure. This study was undertaken to identify the responses of the conifer Cryptomeria japonica, endemic to the Japanese Archipelago, to past climatic changes using a combination of phylogeography and species distribution modelling (SDM) methods. Specifically, this study focused on the locations of refugia during the last glacial maximum (LGM).

Methods

Genetic diversity and structure were examined using 20 microsatellite markers in 37 populations of C. japonica. The locations of glacial refugia were assessed using STRUCTURE analysis, and potential habitats under current and past climate conditions were predicted using SDM. The process of genetic divergence was also examined using the approximate Bayesian computation procedure (ABC) in DIY ABC to test the divergence time between the gene pools detected by the STRUCTURE analysis.

Key Results

STRUCTURE analysis identified four gene pools: northern Tohoku district; from Chubu to Chugoku district; from Tohoku to Shikoku district on the Pacific Ocean side of the Archipelago; and Yakushima Island. DIY ABC analysis indicated that the four gene pools diverged at the same time before the LGM. SDM also indicated potential northern cryptic refugia.

Conclusions

The combined evidence from microsatellites and SDM clearly indicates that climatic changes have shaped the genetic structure of C. japonica. The gene pool detected in northern Tohoku district is likely to have been established by cryptic northern refugia on the coast of the Japan Sea to the west of the Archipelago. The gene pool in Yakushima Island can probably be explained simply by long-term isolation from the other gene pools since the LGM. These results are supported by those of SDM and the predicted divergence time determined using ABC analysis.

Keywords: Approximate Bayesian computation, cryptic refugia, Cryptomeria japonica, genetic divergence time estimation, last glacial maximum, microrefugia, species distribution modelling, sugi, Taxodiaceae

INTRODUCTION

Species distribution shifts and local adaptation through global climatic changes are important factors determining the genetic structure of temperate forest species (Savolainen et al., 2004). Drastic environmental changes have occurred since the last glacial maximum (LGM) around 21 000–18 000 years ago, and most areas of temperate forest vegetation are believed to have been established since the onset of favourable climates through expansion from refugial populations in which they survived the cold and harsh LGM conditions (Tsukada, 1982a; Petit et al., 2003). The contraction and extension of populations may have caused reductions in genetic diversity in the colonizing populations and increases in genetic differentiation among them (Petit et al., 2003), modulated to varying degrees by inputs from secondary contacts (Widmer and Lexer, 2001). Therefore, information on locations of refugia is particularly relevant, given the need to predict the effects of the present period of global warming on species distributions. In many studies using fossil pollen data, a few distinct refugia were shown to be located in regions that escaped the major climatic changes of the glaciations, usually at lower latitudes (Huntley and Birks, 1983; MacDonald, 1993; McLachlan and Clark, 2004). However, the information on fossil pollen is limited to certain areas because fossil pollen is detected mainly in lake sediment or peat deposits. In addition, most perennial plants possess some capacity for vegetative propagation, and clonal reproduction is favoured under certain ecological conditions (Eriksson, 1989, 1996; Pornon et al., 1997, 2000; Dorken and Eckert, 2001). Especially under severe conditions, as in peripheral populations of some plant species, vegetative propagation would be the only means of survival, with little or no flowering (Moriguchi et al., 2001; Eckert, 2002). In this case, the amount of pollen production is low and fossil pollen may not be detected.

Recently, phylogeographic studies have provided support for the existence of previously unknown refugia at higher latitudes (Parducci et al., 2012). The locations of these refugia often conflict with interpretations of pollen evidence and predictions of biogeographical histories of species based on current species distribution patterns. Such refugia are sometimes referred to as ‘cryptic’ (Stewart and Lister, 2001; Provan and Bennett, 2008; Opgenoorth et al., 2010). Species distribution modelling (SDM; Elith and Leathwick, 2009) is a powerful approach to predict past distribution ranges of plant species. In SDM with collated modern distribution data and environmental variables, historical species distributions can be predicted using simulated past climate data (Maiorano et al., 2012). Although several advantages of using SDM have been mentioned, they should be used with caution as they also introduce uncertainty due to biotic interactions, evolutionary change and species dispersal (e.g. Pearson and Dawson, 2003). Our understanding of past processes that have shaped population genetic structures will be improved by combining SDM and phylogeographic data. However, few studies have used this combined approach to predict past (e.g. LGM) distributions of climatically suitable areas or to reconstruct locations of glacial refugia and the dynamics of post-glacial migrations, especially for East Asian temperate forest trees (Sakaguchi et al., 2010, 2011; Qi et al., 2012; Worth et al., 2013).

The Japanese Archipelago extends in a narrow arc from north-east to south-west, with various mountain ranges that mostly run parallel to the arc and probably act as physical barriers to the migration or gene flow of many plant species (Tsukada, 1980). Temperate plant species have generally migrated along the Pacific Ocean side, the Japan Sea side or the mountain slopes of the archipelago with past climate change, expanding from the coasts or to lower altitude refugia to northward or to higher altitudes during interglacial periods (e.g. Cryptomeria japonica, Tsukada, 1980; Fagus, Tsukada, 1982b). Therefore, knowledge of the population history of temperate plant species, including shifts in the distribution, fragmentation and isolation, especially after the LGM, facilitates attempts to interpret their current genetic diversity and structure.

Cryptomeria japonica is an endemic monoecious conifer of Japan, with an outcrossing breeding system involving wind pollination and the production of wind-dispersed seeds. This species is the most economically important forestry tree in Japan. Natural forests of the species are distributed under variable environmental conditions from Aomori Prefecture (40°42′N) to Yakushima Island (30°15′N) in the Japanese Archipelago (Tsumura et al., 2012). Fossil pollen of Cryptomeria was abundant during the interglacial warm intervals in the late Quaternary, reaching a maximum approx. 350 000–380 000 years ago (Tsukada, 1982a). This was one of the most representative species throughout the Quaternary. Cryptomeria japonica was distributed widely over the Japanese Archipelago until 100 000 years ago, and its distribution gradually decreased from 70 000 years ago to 30 000 years ago (Takahara, 1998). Fossil pollen data of this species have been accumulated from many sites, allowing identification of its potential refugia during the LGM and the most likely routes of post-glacial recolonization (Tsukada, 1982a; Miyake et al., 2011). These studies indicated that C. japonica had multiple refugia in Japan during the last glacial period (approx. 18 000 years ago), including the Izu Peninsula, the margins of Wakasa Bay, Oki Island, Muroto Peninsula, Yakushima Island and probably around Kii Peninsula and Toyama Bay (Fig. 1). No clear evidence of northern refugia exists on the main island of Honshu. If northern populations were established by expansion from these refugia, C. japonica would have had to move approx. 130 m per year (Tsukada, 1980). As this is too fast for this species due to its long regeneration time, cryptic refugia may have occurred in the northern part of Honshu where the boreal conifer forests existed during the LGM (Yoshida and Takeuti, 2009).

An external file that holds a picture, illustration, etc.
Object name is mcu19701.jpg

Locations of the 37 sampled populations of Cryptomeria japonica. The grey area represents the current natural distribution of C. japonica (Hayashi, 1960). The black and grey lines indicate the refugia and probable refugia areas, respectively, during the LGM (Tsukada, 1982a).

The geographical variation between natural forests of C. japonica has been investigated, focusing on morphological traits (e.g. needle length or curvature; Murai, 1947), diterpene components (Yasue et al., 1987) and clonal reproduction (Kimura et al., 2013). Previous genetic studies based on single nucleotide polymorphism (SNP) and cleaved amplified polymorphic sequence (CAPS) markers suggested the existence of two main lines: the Japan Sea side lineage and Pacific Ocean side lineage (Tsumura et al., 2007, 2012). Bayesian clustering analyses also provided clear evidence of genetic divergence among four gene pools using these markers: the populations in the north region; the populations on the Japan Sea side; the populations on the Pacific Ocean side; and the populations on Yakushima Island. Despite extensive population genetic studies, the historical process of genetic divergence in this species is still unclear.

This study was performed to identify the responses of C. japonica to past climatic changes by a combination of phylogeography and SDM methods. Specifically, this study focused on the locations and numbers of refugia during the LGM. We also examined the process of genetic divergence among populations using nuclear microsatellite loci to determine the divergence times between genetic clusters. Here, we discuss the historical process of species range formation that could explain the distribution of C. japonica in the Japanese Archipelago.

MATERIALS AND METHODS

Investigated populations

We sampled 761 individual trees of Cryptomeria japonica D.Don from 37 populations located in six districts (Tohoku, Chubu, Kinki, Chugoku, Shikoku and Kyushu) throughout the entire geographic range of C. japonica (Fig. 1, Table 1). Most populations were used in previous studies (Table 1: Takahashi et al., 2005; Tsumura et al., 2007: Kimura et al., 2013). Total DNA was extracted from the needle tissue of each ramet using the modified cetyltrimethylammonium bromide (CTAB) method (Tsumura et al., 1995).

Table 1.

Location of the 37 Cryptomeria japonica populations

Population no.PopulationDistrictnLatitudeLongitudeElevation (m)
1AjigasawaTohoku2140·675556140·205278297
2OwaniTohoku1740·460833140·589167218
3MizusawaTohoku1740·079444140·247500176
4TodosadoTohoku1839·946667140·628611877
5OgaTohoku2039·916111139·756944564
6NibetsuTohoku2339·806111140·260000366
7ShizukuishiTohoku2339·680000140·955278198
8FurukawaTohoku2338·868333140·648333779
9YamanouchiTohoku2238·748333140·074444140
10IshinomakiTohoku2238·328611141·491944159
11*SadoChubu2238·239963138·458551585
12DondenChubu2338·139722138·383333790
13HonnaTohoku2037·426944139·508889415
14*HayatsukiChubu2136·643056137·558889923
15TateyamaChubu2136·583333137·4622221007
16*BijodairaChubu2036·576111137·458889700
17*ObaraChubu1936·519533137·344900917
18KawaduChubu1534·831389139·000000634
19AshitakaChubu2335·232222138·836111805
20FujiChubu935·485556138·612222903
21OwaseKinki1934·344722136·222222918
22AshuKinki2335·307778135·773889886
23ShinguKinki2333·890000135·710000583
24YanaseShikoku2333·592778134·095833762
25AkiShikoku1733·565278133·954444299
26TsuyamaChugoku2035·293056133·967778691
27*ShinjoChugoku2135·224300133·518950805
28OkiChugoku2036·268333133·329167397
29SanbeChugoku2335·130278132·608889567
30WakasugiChugoku2334·695979132·080812468
31*Yoshiwa1Chugoku2234·512317132·068183922
32*Yoshiwa2Chugoku2234·505433132·051633979
33AzoujiChugoku2234·481963131·9633571060
34ShirataniKyushu2330·378056130·573333795
35*YakusugilandKyushu2330·303491130·5730601047
36KuromiKyushu1630·316944130·5102781767
37HanayamaKyushu2230·326667130·4600001247
Average20·6

n, sample size.

*The population was added to this study from previous studies (Takahashi et al., 2005; Tsumura et al., 2007).

Microsatellite analysis

The DNA of organelles with maternal inheritance has been used for phylogeography studies in many plant and animal species because of the clear genetic divergence between populations (McCauley, 1995). However, in this study, we used nuclear DNA markers to examine the historical genetic structure because organelle DNAs such as plastid and mitochondrial DNA show paternal inheritance in this species (Ohba et al., 1971; Kondo et al., 1998). We determined the genotypes of all samples using 20 microsatellite markers, including eight genomic (Moriguchi et al., 2003; Tani et al., 2004) and 12 EST (expressed sequence tag) microsatellite loci (Ueno et al., 2012). The genomic microsatellite loci were also used in previous studies (Takahashi et al., 2005; Kimura et al., 2013) and the EST microsatellite loci have been newly included in this analysis. Polymerase chain reaction (PCR) amplification was performed in 6 μL reaction volumes containing about 5 ng of genomic DNA solution, 3 μL of 2× Qiagen Multiplex Master Mix (Qiagen, Hilden, Germany) and each primer at a concentration of 2 μm (the forward primer was labelled with dye in each primer pair). Amplification was performed using a GeneAmp PCR System Model 9700 thermal cycler (PE Applied Biosystems, Warrington, UK). PCR conditions were as follows: denaturation at 95 °C for 15 min followed by 35 cycles consisting of 94 °C for 30 s, annealing at 57 °C for 90 s and 72 °C for 60 s, with a final extension step at 60 °C for 30 min. The PCR products and DNA size marker (LIZ600; Life Technologies, Foster City, CA, USA) were separated by electrophoresis on an ABI 3100 Avant Genetic Analyser (Applied Biosystems), and their fragment sizes were determined using GENEMAPPER version 3.7 analysis software (Applied Biosystems).

Genetic diversity and characteristics

To assess genetic diversity across all populations, the total number of alleles detected (TA), gene diversity in the total population (HT), average gene diversity within populations (HS; Nei, 1987) and observed heterozygosity (HO) were calculated across all populations at each locus and over all loci using FSTAT v. 2.9.3.2 (Goudet, 2002). The fixation index (FIS; Weir and Cockerham, 1984) was also calculated using FSTAT v. 2.9.3.2 (Goudet, 2002) across all populations at each locus and across all loci to measure departure from Hardy–Weinberg equilibrium (HWE).

Similarly, to assess genetic diversity within each population, the number of alleles (A), allelic richness (Ar; El Mousadik and Petit, 1996), HO and unbiased expected heterozygosity (Nei, 1987) were calculated at each locus and across all loci in each population using FSTAT v. 2.9.3.2 (Goudet, 2002). Ar was calculated for a standard sample size of nine (18 gene copies), which was the smallest sample size, with complete genotypes at all 20 loci, among populations. We also calculated the number of rare alleles (RA; defined as alleles with a frequency <1 % in the total population) and private alleles (PA; unique to one population) per individual in each population. The fixation index (FIS; 1 – HO/HE) was calculated at each locus and over all loci in each population to measure departure from HWE. The deviations of FIS from zero were tested in each population by false discovery rate test using the ‘stats’ package in R 2.15.0 (R Development Core Team, 2012).

We also evaluated whether the populations studied had experienced a reduction in effective population size through founding events or population bottlenecks using Wilcoxon's signed-rank test (one-tailed) with the null hypothesis, HE < HEQ, and the alternative hypothesis, HE > HEQ, under the assumption of mutation–drift equilibrium in the infinite allele model (IAM) and the stepwise mutation model (SMM) with 1000 simulation iterations using BOTTLENECK v. 1.2.02 (Piry et al., 1999).

Bayesian genetic clustering

We performed Bayesian cluster analysis using STRUCTURE 2.3.1 (Pritchard et al., 2000; Falush et al., 2003a) to assign individuals to clusters based on their multilocus genotypes. We first estimated the optimal number of genetic clusters (K) using 761 individuals from 37 natural populations of C. japonica. Using an admixture model with correlated allele frequencies and location information (LOCPRIOR model), 20 independent runs were conducted for each K (putative cluster number, from 1 to 7) with 100 000 iterations after a 10 000-step burn-in. All other parameters were set at their default values. The optimal number of clusters, K, was identified using lnPr (X|K) (Pritchard et al., 2000) or delta K; the rate of change of lnPr (X|K) between successive K values (Evanno et al., 2005) was maximal. The genetic relationships among the genetic clusters were evaluated based on net nucleotide distance calculated in STRUCTURE, and a Neighbor–Joining (NJ) tree based on the net nucleotide distances was generated using POPULATIONS v. 1.2.30 (Langella, 1999). The net nucleotide distance is the average probability that the two alleles in a pair from two different genetic clusters are different (Falush et al., 2003b).

CLUMPP1.1.1 (Jakobsson and Rosenberg, 2007) was used to calculate the average membership coefficient (Q value) for each individual by aligning and converging the results of the 20 runs described above. We also calculated the average membership coefficient (Q value) for each individual excluding the Yakushima Island population (K = 3) for ABC analyses (see below) because the genetic cluster of Yakushima Island was particularly different from the other gene pools. In addition, principal co-ordinate analysis (PCoA) we conducted based on genetic distance in GeneAlEx 6·41 (Peakall and Smouse, 2006) to visualize the relationships among populations.

Approximate Bayesian computation (DIY ABC)

In the absence of a fossil record, the time scales which estimated solely on genetic data may not be as precise as geological time scales, but are good enough to make it possible to generate the temporal hierarchies required for understanding interspecific gene flow and demographic dynamics cause by distributional changes (Wakeley, 2008). For probabilistic analysis among alternative hypotheses for the history of gene pool divergence, we used the approximate Bayesian computation procedure (Beaumont et al., 2002) in DIY ABC v. 1.0.4.39 (Cornuet et al., 2008, 2010). We tested the following six competing scenarios (for details, see Fig. 2 and Supplementary Data Fig. S1).

An external file that holds a picture, illustration, etc.
Object name is mcu19702.jpg

Distribution of cluster memberships at individual and population levels in the 37 Cryptomeria japonica populations obtained by STRUCTURE analysis (Pritchard et al., 2000). Number of cluster (K) is 4. Each horizontal bar in the histograms represents the proportion of cluster memberships for each individual. The pie charts superimposed on the maps show the average proportions of cluster memberships among the individuals sampled in each population.

Scenario 1 (three types of refugia)

Gene pools B2, B3 and B4 diverged at the same time, and then gene pool B1 arose recently via divergence from gene pool B2. This scenario suggests that three types of refugia (e.g. the margins of Wakasa Bay, the Izu Peninsula and Yakushima Island as sampled sites of fossil pollen) existed at the LGM and that the population in northern Tohoku (B1) diverged when this species expanded its distribution to the north.

Scenario 2 (four types of refugia)

Four gene pools diverged at the same time. This scenario suggests that several refugia existed, including cryptic refugia, in the northern Tohoku region at the LGM.

Scenario 3 (two lineages diverged before the LGM)

First, the populations on the Japan Sea side (B1 and B2) and Pacific Ocean side (B3 and B4) diverged before the LGM, and each gene pool then separated.

Scenario 4 (divergence and admixture)

Scenario 4 includes one admixture event because most populations showed admixed gene pools on the Pacific Ocean side, except the Yakushima populations, in clustering K = 2 (see the Results and Supplementary Data Fig. S2). First, the populations on the Japan Sea side (B1 and B2) and Pacific Ocean side (B4) diverged, and gene pool B3 arose via admixture between Japan Sea side and Pacific Ocean side populations. Finally, gene pool B1 arose recently via divergence from gene pool B2.

Scenario 5 (divergence three times)

First, gene pool B4 diverged and then gene pool B3 separated. Finally, gene pool B1 arose recently via divergence from gene pool B2. Scenario 5 reflects the net nucleotide distance among the four gene pools.

Scenario 6

First, we incorporated a bottleneck into the history of gene pool B1 because northern populations may have suffered from small population size in the glacial period. However, the time of the bottleneck in gene pool B1 could not be estimated well in the best model (Scenario 2), and we also tested the best model without the bottleneck as Scenario 6.

In this study, we conducted a modified version of FST outlier loci analysis (Beaumont and Nichols, 1996) under the assumption of the IAM using LOSITAN (Antao et al., 2008). Simulations were run 10 000 times to generate 95 % confidence intervals (CIs). The default value for the false discovery ratio (= 0·1) was used in this analysis. ABC analyses were applied to data with 14 loci, excluding six outlier loci (Supplementary Data Table S1). Because it is difficult to decide the gene pool in high admixed genets, 233 individuals with Q values >0·8 were used as samples for ABC analyses. From Yakushima Island in Kyushu district, 34 individuals were extracted as samples of gene pool B4 in clustering K = 2. In total, 199 individuals with Q values >0·8 from multiple populations were used as samples for the other three gene pools (B1, B2 and B3) in the clustering K = 3, excluding the Yakushima Island population.

In DIY ABC, we employed the higher mutation rate from the generalized stepwise mutation model (GSM; Estoup et al., 2002) and the lower rate for single nucleotide indels (SNIs) as mutation models of simple sequence repeats (SSRs). The default prior values were used for all of these parameters (Supplementary Data Table S3). Summary statistics were calculated in each scenario for selecting scenario estimating posteriors of demographic parameters (Table S4). In total, 4 million data sets were simulated, of which the 1 % closest to the observed data were used to estimate the relative posterior probabilities of each scenario via a logistic regression approach. Equal prior probability was set for each scenario. The relative likelihoods of the four scenarios were compared by logistic regression on 1 % of the closet simulated data sets.

Current and past distribution modelling

To assess the effects of the last glacial distribution shifts on the population demography of C. japonica, the potential habitats present at the LGM (21 000 years bp) and under current climate conditions were derived by SDM. The generalized additive model (GAM) (Elith et al., 2006) was allocated based on the presence and absence data using the ‘mgcv’ package in R 2.15.0 (R Development Core Team, 2012). The distribution map of C. japonica natural forests by Hayashi (1960) was used to obtain data for the presence of this species. The area without natural forests of C. japonica was used as data for absence of this species. Four climatic variables (Bio6, mean daily minimum temperature of winter; Bio10, mean temperature of summer; Bio18, summer precipitation; and Bio19, winter precipitation) at a spatial resolution of 2·5 min were used as explanatory variables when extracting data from the WorldClim database (http://www.worldclim.org). The four climatic variables that are generally thought to be important for plant growth and survival were used as explanatory variables in the SDM. These variables have been used to obtain highly accurate SDMs in East Asia (e.g. Tanaka et al., 2012; Tsuyama et al., 2014; Yun et al., 2014). The dataset is based on meteorological observations collected for the period 1950–2000 (Hijmans et al., 2005). The area under the curve (AUC) value, which was derived from receiver operating characteristic (ROC) analysis (Metz, 1978; Thuiller et al., 2003), was used to assess the prediction accuracy of the GAM. The AUC value was calculated by comparing the predicted presence/absence data and the validation data bootstrapped from all of the training data with 100 repetitions.

The relative importance of each variable was used to measure the magnitude of the contribution of each explanatory variable in the model, which was defined as the sum of the deviance of variables across all models in which the variables occurred (Burnham and Anderson, 2002). The coefficient values were also calculated for four current climatic variables at each mesh.

We projected the potential habitat for C. japonica using GAM on the climate of the LGM as a representative of glacial periods during the Pleistocene. To project the LGM distribution of C. japonica, we applied two GCMs for the LGM (approx. 21 000 years ago): the Community Climate System Model version 3 (hereafter CCSM) (Collins et al., 2006) and the Model for Interdisciplinary Research on Climate 3.2_hires (hereafter MIROC) (K-1 Model Developers, 2004). We downloaded the data from the WorldClim dataset (http://www.worldclim.org/), which were spatially interpolated to 2·5 min. To validate the relevance of projected LGM occurrences for C. japonica, we calculated minimum distances between fossil occurrence and projected potential habitat for the species, and tested the difference between them by sensitivity analysis. Calculating the distances between C. japonica pollen data and the projected potential habitats for the LGM, the distances between the pollen presence and the predicted presence under the MIROC simulation were 14·8 ± 23·87 km, and the sensitivity was 72·7 % (Supplementary Data Fig. S3A). In the case of CCSM simulation, the minimum distance and sensitivity were 185·9 ± 179·93 km and 0·9 %, respectively (Fig. S3B). From this result, the MIROC simulation was used in this study.

RESULTS

Genetic diversity across all populations and within populations

The 20 microsatellite loci used were highly polymorphic. The total number of alleles detected over all populations at each locus ranged from two to 46, with an average value of 15·6 (Supplementary Data Table S2). On average, across all loci, the gene diversity in the total population (HT) and the average gene diversity within populations (HS) were 0·651 and 0·628, respectively. Genomic and EST SSR markers that showed high degrees of polymorphism were used to examine the diversity and structure in this species, in which the ranges of gene diversity in the total population (HT) were 0·661–0·946 and 0·157–0·884, respectively (Table S2).

High levels of genetic diversity within populations were also observed in each population (on average, A = 6·6, Ar = 5·4, RA = 0·74, HO = 0·592, HE = 0·611; Supplementary Data Table S2). Rare alleles (RA) were found in all populations. Private alleles (PA) were found in 21 of 37 populations and heterozygosity excess was detected in four populations (one Tohoku population, two Chugoku populations and one Kyushu population) after Bonferroni correction under the IAM (Table S2). The FIS values over all loci significantly positively deviated from zero in nine populations (four Tohoku populations, one Kinki population, one Shikoku population and three Chugoku populations; Table S2).

Genetic structure

The population genetic structure was explored by the STRUCTURE analysis. The values of log-likelihood of the data, ln P(X|K), reached a plateau when the numbers of clusters were K > 4 (Supplementary Data Fig. S4). Furthermore, the values of delta K, based on the rate of change of ln P(X|K) between successive K values, were highest when K was 2. Therefore, we examined the proportional membership of each cluster of each individual and population at K = 4 and K = 2. When we set K to 4, the distribution of cluster membership exhibited not only the pattern of population divergence between the Japan Sea and Pacific Ocean sides, but also genetic sub-structures among populations along each side (Fig. 2). On the Japan Sea side, the membership proportion of gene pool B1 was higher in the populations from northern Tohoku district (population nos 1–7) to Sado Island in Niigata Prefecture (nos 11, 12), whereas that of gene pool B2 was higher in the populations (nos 17, 26–33) from the Chubu to Chugoku districts (Fig. 2). Similarly, among the populations along the Pacific Ocean side, the populations (nos 10, 18–25) from Tohoku to Shikoku district had higher proportions of gene pool B3, whereas the Yakushima Island populations (nos 34–37) in Kyushu district had higher proportions of gene pool B4. High admixture of several gene pools was found in Kyoto, Toyama Prefecture and southern Tohoku district (nos 8, 9, 14–16, 22). The F-values of gene pools B1, B2, B3 and B4 were 0·05937, 0·04393, 0·02779 and 0·05026, respectively (Fig. 2). The NJ tree using net nucleotide distance showed that gene pool B4 was markedly divergent from the other three gene pools (Fig. 2). The distribution of membership of the two clusters (A1 and A2) was found to be geographically structured between the populations along the Japan Sea and Pacific Ocean sides (Supplementary Data Fig. S2). The proportion of gene pool A1 (blue) was higher in the populations of northern Tohoku district, whereas that of gene pool A2 (red) was higher in populations of Yakushima Island in the Kyushu district. The F-values of gene pools A1 and A2 were 0·02810 and 0·03779, respectively (Fig. S2).

This result of population genetic structure by STRUCTURE analysis corresponded to that of PCoA analysis (Fig. 3). The first two axes accounted for 65·6 % of the total variation. The first axis reflected Yakushima Island, other Pacific Ocean side populations and Japan Sea side populations. The second axis reflected northern Tohoku populations and other populations on the Japan Sea side (Fig. 3).

An external file that holds a picture, illustration, etc.
Object name is mcu19703.jpg

Genetic similarity among 37 populations, as depicted from principal co-ordinate analysis. Pie charts represent the proportion of cluster memberships in each population using STRUCTURE analysis with K = 4 (see Fig. 2). Gene pool B1 = blue, gene pool B2 = sky blue, gene pool B3 = orange, gene pool B4 = red.

ABC analysis

A comparison of posterior probabilities of the six scenarios using local linear regression indicated that Scenario 6 had the highest support, with a probability of 0·811 (Table 2). Scenario 6 had the smallest number of marginal summary statistics (Supplementary Data Table S4). ABC analysis provided good support for Scenario 6, with the same time divergence of four gene pools and no bottleneck (Supplementary Data Fig. S1, the type I error rate = 0·22).

Table 2.

Posterior probability of each scenario and its 95 % confidence interval based on the logistic estimate according to DIY ABC

SimulationScenarioPosterior probability95 % CI (lower–upper)
First set10·1800·138–0·022
20·9340·924–0·945
30·0050·004–0·007
40·0220·017–0·027
50·0200·016–0·025
Second set20·1890·158–0·221
60·8110·779–0·842

The median value of the divergence time (t2) was estimated as 507 generations (Supplementary Data Figs S1 and S5; 95 % CI = 161–1170). For Scenario 6, the median values of effective population size of NB1, NB2, NB3, NB4 and NA before population divergence were 4520, 8040, 8840, 5330 and 2630, respectively (Table 3). Posterior distributions are shown in full in Table 3.

Table 3.

Demographic parameters obtained by DIY ABC

Parameter
MeanMedianModeQuantile 2·5 %Quantile 5·0 %Quantile 95·0 %Quantile 97·5 %
Effective population sizeNB14760452037701380173086609240
NB27780804086204350499097109860
NB38590884091005800649098409920
NB45420533049001780221090609510
NA29802630156029050567907750
Time scale in generationst256750745512616111701360
Mean mutation rateMean mutation rate_SSR2·00 × 10–41·60 × 10–41·18 × 10–41·00 × 10–41·04 × 10–44·36 × 10–45·57 × 10–4
Mean P*0·1160·1060·1000·1000·1000·1700·207
Mean mutation rate_SNI8·93 × 10–59·12 × 10–59·34 × 10–56·84 × 10–57·29 × 10–59·91 × 10–59·96 × 10–5

*The parameter of the geometric distribution to generate multiple stepwise mutation.

Current and past (LGM) distributions of C. japonica

For C. japonica, the AUC test (mean ± s.d.) for SDM was high (0·821 ± 0·003). Although the current distributional predictions were good representations of the extant distribution of species (Fig. 4A), we also found some potential habitats where the species does not occur at present, probably due to human activity (e.g. low land on the Japan Sea side; Fig. 4A). In contrast, the distribution model for the LGM (Fig. 4B) differed substantially from the current distribution, indicating a general southward range shift. Predicted suitable habitats for this species at the LGM were restricted to lowland areas, such as Oga Peninsula, Sakata (near pop 9), Toyama Bay, Oki Island, Kii Peninsula and Yakushima Island, indicating an even smaller geographic range than at present (Figs 1 and and44B).

An external file that holds a picture, illustration, etc.
Object name is mcu19704.jpg

Potential distributions as probability of occurrence (A) at present and (B) at the last glacial maximum (LGM, 21 000 years bp) for Cryptomeria japonica.

The relative importance of variables (IOV) indicated that winter precipitation (Bio19, 50·2 %) was the most influential factor for C. japonica, followed by summer precipitation (Bio18, 27·2 %), mean temperature of summer (Bio10, 14·9 %) and the mean daily minimum temperature of winter (Bio6, 7·7 %). The regional variation of coefficient value was different among current climatic variables (Supplementary Data Fig. S6).

DISCUSSION

Genetic structure of C. japonica

The results of PCoA analysis and Bayesian clustering analysis indicated clear genetic divergence of two gene pools (A1 and A2; Supplementary Data Fig. S2) and four gene pools (B1–B4; Fig. 2). Using the delta K (K = 2) indicates two major gene pools of C. japonica: one distributed along the Japan Sea side (Japan Sea lineage, A1) and one along the Pacific Ocean side (Pacific Ocean lineage, A2). Our results were consistent with previous studies using CAPS markers (Tsumura et al., 2007) and SNP markers (Tsumura et al., 2012). Using 148 CAPS markers, Bayesian analysis of the genetic structure defined eight groups for 29 natural populations (Tsumura et al., 2007). Tsumura et al. (2012) also found seven groups (highest log-likelihood K) and two groups (delta K) for 14 natural populations, according to structure analysis based on results from 1023 SNPs, which provided a much clearer genetic structure than that from 148 CAPS markers. In this study, we used only 20 loci but highly polymorphic markers with a total of 311 alleles corresponding to those of Tsumura et al. (2007) in terms of the number of alleles, which may be sufficient to detect the genetic structure in this species.

The genetic structures show that two major groups (i.e. the Pacific Ocean side and the Japan Sea side) have also been reported in other tree species (e.g. Fagus crenata, Hiraoka and Tomaru, 2009; Kalopanax septemlobus, Sakaguchi et al., 2011). The Japan Sea side of the country has heavy snowfall during the winter season, whereas the Pacific Ocean side is quite dry during the winter. The contrasting climatic conditions, especially the difference in amount of snowfall, have strongly influenced the characteristics of tree species, resulting in differences between the two sides of the country. Geographical variation in C. japonica was also observed in morphological traits (Murai, 1947), diterpene components (Yasue et al., 1987) and clonal reproduction (Kimura et al., 2013) between the Japan Sea side and Pacific Ocean side. Each gene pool has probably been adapted to local environmental conditions, leading to the current genetic structure of this species, including neutral and adaptive divergence between populations. In a previous study, several candidate SNPs for local adaptation were detected, especially between populations of the Japan Sea side and Pacific Ocean side (Tsumura et al., 2012).

Prediction of divergence time using ABC analysis

The ABC analysis using 14 SSR loci supported Scenario 6, in which four gene pools separated at the same time, rather than gradual divergence scenarios. The time of divergence (t2) was predicted as 507 generations. These results suggested that four ancestral populations were present before approx. 500 generations. This divergence time is dependent on the generation time of this species, but the generation time of woody species can be much longer than their maturation age (Petit and Hampe, 2006). As C. japonica begins to produce cones at 20 years of age in artificial open habitats (e.g. Endo et al., 2012), we therefore assumed a generation time for this species of approx. 150 years under natural conditions. If we assume the generation time to be 150 ± 50 years, the divergence time for the four gene pools could be 76 050 ± 25 305 years bp. Although the estimated divergence time also changes by the mutation rate, the posterior distribution of model parameters under the most likely scenario was used to make inferences about the timing of divergence before the LGM. The four gene pools of C. japonica probably indicated refugia populations at the LGM.

Climatic conditions controlling C. japonica distribution

The distribution of the potential habitats for C. japonica was predicted on the basis of the GAM using the present distribution data of C. japonica and a current climatic database. The prediction accuracy of the potential habitats was evaluated as ‘good’ based on the AUC. Species distribution modelling showed that the ecological niche of C. japonica is constrained primarily by winter precipitation (Bio19, 50·2 %). The distribution probability was low in Kanto and Hokkaido districts (Fig. 4). In these areas, climatic variables also showed low coefficient values (mean temperature of summer and winter precipitation in Kanto, mean daily minimum temperature of winter and summer precipitation in Hokkaido; Supplementary Data Fig. S6). Tsukada (1982) noted that the probable glacial refugia in Cryptomeria can be characterized by climatic variables: annual precipitation of at least 1200 mm, annual temperature >5·0 °C, minimum mean January temperature of –7·0 to –8·2 °C and mean July temperature of 20·0–25·0 °C. This climatic regime is almost equivalent to that at the northern limit of planted Cryptomeria forest in Hokkaido today (Tsukada, 1980). Our results were consistent with the report by Tsukada (1982). In addition, the current distribution of the predicted potential habitats covered the natural distribution map of C. japonica by Hayashi (1960). In this model, past human activity, such as logging of natural populations, was not considered; therefore, the predicted current potential habitat area would be wider, especially in the lowlands, if we took past human activity into consideration, but no contradiction occurred with the current positions of natural populations.

Comparing pollen data with predicted potential habitats for the LGM, the MIROC simulation was suggested to be more realistic than CCSM in the case of Japan (Supplementary Data Fig.S3). Calculating the distances between C. japonica pollen data and the projected potential habitat during the LGM showed that the projected potential habitat on the MIROC is reasonably accurate. Potential glacial refugia of C. japonica were predicted not only in western Japan, e.g. around Oki Island, Kii Peninsula and Yakushima Island, but also in lowlands of Tohoku district, such as Oga Peninsula. These results suggested the presence of putative refugia in the coastal areas of Tohoku district.

Many suitable habitats for this species were predicted in Kyushu district at the LGM; however, these natural populations were thought to have been destroyed by volcanic eruptions. For example, a large-scale eruption of Mt. Aso occurred 90 000 years ago and the Kikai caldera erupted violently in the southern Kyushu district 7300 years ago (Machida and Arai, 1992), causing catastrophic damage in these areas where plant life was thought to have been virtually annihilated. Although the small old forests of C. japonica distributed on Kyushu Island, this population might be established by the immigration from Chugoku populations, such as Azouji and others, after the extinction of natural forest on Kyushu Island (Takahara, 1998). The supporting result was obtained by the genetic analysis using SNP markers (Tsumura et al., 2012).

Evidence of cryptic refugia of C. japonica in northern Japan

Bayesian clustering analysis with the highest log-likelihood values also divided into four gene pools: northern Tohoku district (B1); from Chubu to Chugoku districts (B2); from Tohoku to Shikoku district on the Pacific Ocean side (B3); and the Yakushima Island population (B4). The results of PCoA analysis also supported these gene pools. Axis 1 (37·7 %) represents genetic differences among Yakushima Island populations, the other Pacific Ocean side populations and Japan Sea side populations, and axis 2 (27·9 %) represents genetic difference between the northern Tohoku and the other Japan Sea side populations (Fig. 3).

Relatively high levels of within-population gene diversity were detected in the populations of Tohoku district (HE: average 0·604, range 0·581–0·642) in comparison with the other districts (HE: average 0·614, range 0·570–0·643). In addition, no evidence of a recent bottleneck was observed in Tohoku district (except the Oga population). These results suggest the potential of northern cryptic refugia and/or the potential of admixture events from several refugia between populations in Tohoku district. The results of STRUCTURE analysis in K = 4 indicated the relatively purity of gene pool B1 in northern Tohoku populations (Fig. 2). We also found that some populations in northern Tohoku possessed private alleles and/or rare alleles. This genetic structure suggested the presence of cryptic refugia in northern Japan at the LGM. Recent phylogeographic studies on temperate trees and animals in Asia suggested that small populations may have survived in the northernmost part or high elevation mountain area and separated from the southern large glacial refugia (Tsuda and Ide, 2005; Tsumura et al., 2007; Ohnishi et al., 2009; Opgenoorth et al., 2010). Tsukada (1982) predicted probable refugia at Toyama Bay using environmental data, such as annual precipitation and temperature, and fossil pollen records. In this case, the C. japonica forest had to expand at a rate of 130 m per year from Toyama Bay (near population no. 14) to Akita Prefecture (near population no. 6) after the LGM to establish the current distribution (Tsukada 1980). In contrast, a few studies found a low frequency (approx. 10 %) of fossil pollen of C. japonica 12 000 years ago at Oga Peninsula (Yoshida and Takeuti, 2009). Moreover the fossil pollen of C. japonica was found in northern Tohoku district at least 5000–6000 years ago (e.g. Shimokita Peninsula; Kawamura, 1979; Yamanaka et al., 1990). These results correspond to the consideration of Takahara (1998) that scattered and limited populations could have survived in geographically isolated pockets of favourable habitat within a forest mosaic dominated by boreal taxa in northern Japan during the LGM. Our SDM also predicted that this species could have been distributed in northern Tohoku district (the coastal side from population no. 9 and Oga Peninsula) during the LGM, although population sizes in these areas were limited. Tsukada (1982) suggested that small isolated refugial populations could hardly persist in the harsh sub-arctic climate, such as the northern region of Japan, during the full glacial period. However, the effective population size of gene pool B1 was sufficient to maintain populations and genetic diversity (median 4650, 95 % CI 1640–8630), although this number was the smallest compared with other gene pools.

Frequent layering from trunks and/or branches occurs in this species, especially in snowy regions (Kimura et al., 2013). As there are large amounts of snowfall over the mountainous regions on the Japan Sea side, the main regeneration system for C. japonica in these areas is generally accepted to be layering, which occurs when a flexible branch or stem is bent to the ground by snow pressure (Shimizu et al., 2002; Hirayama and Sakimoto, 2008). Several studies using genetic markers have shown a higher frequency of layering in regions with heavy snowfall (Taira et al., 1997; Moriguchi et al., 2001; Hirayama and Sakimoto, 2008; Kimura et al., 2013). The reproductive strategy of C. japonica was found to be strongly affected by genetic factors and snow depth (Kimura et al., 2013). These results suggest that clonal reproduction allows persistence under harsh conditions, such as in the LGM, where establishment of small individuals, e.g. seedlings, is disturbed by heavy snow. Clonal reproduction as a reproductive strategy in C. japonica contributes to maintenance of genetic variation by increasing the life span of rare genotypes with unique alleles and by preserving the few seedlings that are able to establish under conditions of frequent disturbance. Such a reproductive strategy may explain the persistence of the small refugial populations in northern Tohoku district.

Potential of anagenetic divergence on Yakushima Island

The results of STRUCTURE analysis in K = 4 also indicated the specificity of gene pool B4 in Yakushima Island populations (Fig. 2). Abundant Cryptomeria pollen occurrence (>40 %) was recorded during the Pliocene on southern Kyushu Island, although sediments derived from the full- and late-glacial periods have yielded neither Cryptomeria pollen nor its microscopic remains, suggesting that it became extinct in Kyushu Island before the LGM (Tsukada, 1980, 1982a). The main island (Honshu), Shikoku and Kyushu of the Japanese Archipelago were connected and formed one large island during the LGM. However, the Kikai caldera near Yakushima Island erupted violently 7300 years ago (Machida and Arai, 1992), causing catastrophic damage to the forests in Kyushu district and Yakushima Island (Kimura et al., 1996). Fossil pollen data indicated that Yakushima Island was one of the main refugia at the LGM (Tsukada, 1982a). Therefore, Yakushima Island may have been isolated from the other C. japonica populations for a long time.

Anagenesis (also known as phyletic speciation) has attracted attention as a speciation process, especially on oceanic islands, in which the initial founder populations simply diverge over time without further specific differentiation (Stuessy et al., 1990, 2006; Whittaker et al., 2008; Takayama et al., 2012; Zou et al., 2013). The clear genetic distinctions and similar genetic diversity were found between common species distributed on the Asian continent and endemic species distributed on islands as evidence of anagenetic speciation (Takayama et al., 2012; Zou et al., 2013). Bayesian clustering analyses showed a relatively high F-value in gene pool B4, suggesting an episode of genetic drift and/or selection during colonization. No clear evidence of a bottleneck was detected based on allelic frequency distribution and excess of observed heterozygosity; indeed, genetic diversity was high compared with other populations. Both PA and RA were also found in Yakushima Island populations. The old stumps have remained in extremely fresh condition in spite of the high levels of precipitation and humidity on the island due to their high resin contents (Toda and Sato, 1969; Takahashi et al., 2008). This result seems to show the difference in wood traits, although the effect of environmental condition was not avoided. Needle shape is also different compared with those of trees from other regions (M. K. Kimura, unpubl. data). The needles of individuals from Yakushima Island are short, opened and hard regardless of planted environment. These results probably suggest an anagenetic process on Yakushima Island. In fact, many endemic species on Yakushima Island are found in several taxa (Yahara et al., 1987).

Admixture-like structure of C. japonica

High admixture of several gene pools was found in Kyoto, Toyama Prefecture and southern Tohoku district (population nos 8, 9, 14–16, 22). This result suggests that multiple refugia and/or admixture events, with an ambiguous genetic cline along latitudes, would be found in C. japonica. We found potential cryptic refugia in Tohoku district (B1), further north than previous pollen records (Tsukada, 1982a). On the main Island of Japan (Honshu), the other two clear gene pools (B2 and B3) were also found from Chubu to Chugoku district on the Japan Sea side and from Tohoku to Shikoku district on the Pacific Ocean side, reflecting a refugial area (Wakasa Bay, Kii Peninsula and Izu Peninsula) based on fossil pollen analysis (Tsukada, 1982a). Species distribution modelling also predicted that these areas would be suitable with high probability for C. japonica at the LGM. It has numerous geographical barriers (e.g. high mountain ranges, valleys and seaways), which are expected to have influenced the distribution patterns and past migration of tree species in Japan. For example, the high mountain range divides the main island of Honshu into the Pacific Ocean and the Japan Sea sides. However, it is thought that secondary contacts could easily have occurred in some areas of lower mountains, because the surroundings of the locations of populations showing admixture have no geographical barrier such as higher mountains.

Evidence of multiple refugia, such as high genetic diversity in marginal populations or contact populations, has been reported in many species (e.g. Hewitt, 2000; Petit et al., 2003). The existence of multiple refugia may have played important roles in maintaining the genetic diversity in these areas. Admixture of genetically differentiated populations would result in high genetic diversity in the newly established populations comparable with other populations. In admixed zones, elevated genetic diversity has been reported in several tree species, e.g. Fagus sylvatica, Fraxinus excelsior, Pinus resinosa and Kalopanax septemlobus (Comps et al., 2001; Walter and Epperson, 2001; Petit et al., 2003; Heuertz et al., 2004; Sakaguchi et al., 2011).

Conclusions

The combined evidence from nuclear microsatellites and SDM clearly indicated that climate-driven range fluctuations have shaped population genetic diversity and structure of C. japonica in the Japanese Archipelago. The gene pool detected in northern Tohoku district is likely to have been generated by cryptic northern refugia in the coastal landscapes on the Japan Sea side. The characteristic gene pool on Yakushima Island could probably be explained simply by isolation from the other gene pools since the LGM, e.g. by anagenetic divergence processes. An ambiguous genetic diversity cline along elevations would be predicted by the distribution pattern and accessibility of refugial populations after the LGM, suggesting that multiple refugia and/or admixture events affected the overall structure of genetic diversity. These results were consistent with those of SDM and predicted divergence time using ABC analysis.

ACKNOWLEDGEMENTS

The authors are grateful to Dr T. Nagamitsu for valuable discussions. We also thank Y. Taguchi, M. Koshiba and Y. Komatsu for their help in DNA analyses. This study was partly supported by a grant from the Forest Agency of Ministry of Agriculture, Forestry and Fishery, and by the Programme for the Promotion of Basic and Applied Research for Innovations in Bio-oriented Industry. M.K.K., K.U., Y.M. and Y.T. conceived and designed the experiments. M.K.K. and L.S.J.M. performed molecular analyses and statistical analyses. K.N. estimated current and past distributions by species distribution modelling. M.K.K. and Y.T. wrote the manuscript; the other authors provided editorial advice.

LITERATURE CITED

  • Antao T, Lopes A, Lopes R, Beja-Pereira A, Luikart G. LOSITAN: a workbench to detect molecular adaptation based on a Fst-outlier method. BMC Bioinformatics. 2008;9:323. [Europe PMC free article] [Abstract] [Google Scholar]
  • Beaumont MA, Nichols RA. Evaluating loci for use in the genetic analysis of population structure. Proceedings of the Royal Society B: Biological Sciences. 1996;263:1619–1626. [Google Scholar]
  • Beaumont MA, Zhang W, Balding DJ. Approximate Bayesian computation in population genetics. Genetics. 2002;162:2025–2035. [Europe PMC free article] [Abstract] [Google Scholar]
  • Burnham KP, Anderson DR. Model selection and multi-model inference: a practical information-theoretic approach. Berlin: Springer; 2002. [Google Scholar]
  • Collins WD, Bitz CM, Blackmon ML, et al. The Community Climate System Model Version 3 (CCSM3) Journal of Climate. 2006;19:2122–2143. [Google Scholar]
  • Comps B, Gömöry D, Letouzey J, Thiébaut B, Petit R. Diverging trends between heterozygosity and allelic richness during postglacial colonization in the European beech. Genetics. 2001;157:389–397. [Europe PMC free article] [Abstract] [Google Scholar]
  • Cornuet J-M, Santos F, Beaumont MA, et al. Inferring population history with DIY ABC: a user-friendly approach to approximate Bayesian computation. Bioinformatics. 2008;24:2713–2719. [Europe PMC free article] [Abstract] [Google Scholar]
  • Cornuet J-M, Ravigné V, Estoup A. Inference on population history and model checking using DNA sequence and microsatellite data with the software DIYABC (v1. 0) BMC Bioinformatics. 2010;11:401. [Europe PMC free article] [Abstract] [Google Scholar]
  • Dorken M, Eckert C. Severely reduced sexual reproduction in northern populations of a clonal plant, Decodon verticillatus (Lythraceae) Journal of Ecology. 2001;89:339–350. [Google Scholar]
  • Eckert C. The loss of sex in clonal plants. Evolutionary Ecology. 2002;15:501–520. [Google Scholar]
  • Elith J, Leathwick JR. Species distribution models: ecological explanation and prediction across space and time. Annual Review of Ecology, Evolution, and Systematics. 2009;40:677–697. [Google Scholar]
  • Elith J, Graham C, Anderson R, et al. Novel methods improve prediction of species' distributions from occurrence data. Ecography. 2006;29:129–151. [Google Scholar]
  • El Mousadik A, Petit RJ. High level of genetic differentiation for allelic richness among populations of the argan tree [Argania spinosa (L.) Skeels] endemic to Morocco. Theoretical and Applied Genetics. 1996;92:832–839. [Abstract] [Google Scholar]
  • Endo R, Fukushima S, Takeyama F, Kobayashi S, Iwasawa M. Change of mast year male flower setting of Japan cedar (Cryptomeria japonica D. Don) plus tree clones accompanying aging from 26 to 42-years-old (subtitle in Japanese) Journal of the Japanese Forest Society. 2012;94:252–256. [Google Scholar]
  • Eriksson O. Seedling dynamics and life histories in clonal plants. Oikos. 1989;55:231–238. [Google Scholar]
  • Eriksson O. Regional dynamics of plants: a review of evidence for remnant, source–sink and metapopulations. Oikos. 1996;77:248–258. [Google Scholar]
  • Estoup A, Jarne P, Cornuet JM. Homoplasy and mutation model at microsatellite loci and their consequences for population genetics analysis. Molecular Ecology. 2002;11:1591–1604. [Abstract] [Google Scholar]
  • Evanno G, Regnaut S, Goudet J. Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Molecular Ecology. 2005;14:2611–2620. [Abstract] [Google Scholar]
  • Falush D, Stephens M, Pritchard JK. Inference of population structure using multilocus genotype data: linked loci and correlated allele frequencies. Genetics. 2003a;164:1567–1587. [Europe PMC free article] [Abstract] [Google Scholar]
  • Falush D, Wirth T, Linz B, et al. Traces of human migrations in Helicobacter pylori populations. Science. 2003b;299:1582–1585. [Abstract] [Google Scholar]
  • Goudet J. FSTAT, a program to estimate and test gene diversities and fixation indices (version 2.9.3.2) 2002. Available from http://www2.unil.ch/popgen/softwares/fstat.htm .
  • Hayashi Y. Taxonomical and phytogeographical study of Japanese conifers. Tokyo: Norin-Shuppan; 1960. [Google Scholar]
  • Heuertz M, Hausman JF, Hardy OJ, Vendramin GG, Frascaria-Lacoste N, Vekemans X. Nuclear microsatellites reveal contrasting patterns of genetic structure between western and southeastern European populations of the common ash (Fraxinus excelsior L.) Evolution. 2004;58:976–988. [Abstract] [Google Scholar]
  • Hewitt G. The genetic legacy of the Quaternary ice ages. Nature. 2000;405:907–913. [Abstract] [Google Scholar]
  • Hijmans R, Cameron S, Parra J, Jones P, Jarvis A. WorldClim, version 1.3. Berkeley: University of California; 2005. http://biogeo.berkeley.edu/worldclim/worldclim.htm . [Google Scholar]
  • Hiraoka K, Tomaru N. Genetic divergence in nuclear genomes between populations of Fagus crenata along the Japan Sea and Pacific sides of Japan. Journal of Plant Research. 2009;122:269–282. [Abstract] [Google Scholar]
  • Hirayama K, Sakimoto M. Clonal structure and diversity of Cryptomeria japonica along a slope in a cool-temperate, old-growth mixed forest in the snowy region of Japan. Canadian Journal of Forest Research-Revue Canadienne de Recherche Forestière. 2008;38:2804–2813. [Google Scholar]
  • Huntley B, Birks HJB. An atlas of past and present pollen maps for Europe: 0–13,000 years ago. Cambridge: Cambridge University Press; 1983. [Google Scholar]
  • Jakobsson M, Rosenberg NA. CLUMPP: a cluster matching and permutation program for dealing with label switching and multimodality in analysis of population structure. Bioinformatics. 2007;23:1801–1806. [Abstract] [Google Scholar]
  • K-1 Model Developers. K-1 coupled GCM (MIROC) description. K-1 Technical Report. 2004;1:1–34. [Google Scholar]
  • Kawamura T. Special reference to the distribution of Cryptomeria japonica. Quaternary Research. 1979;18:79–88. [Google Scholar]
  • Kimura K, Ooi N, Suzuki S. Evidence of vegetation recovery on Yakushima Island after the major Holocene eruption at the Kikai caldera, as revealed by the pollen record of buried soils under the old-growth Cryptomeria japonica forest. Japanese Journal of Historical Botany. 1996;4:13–23. [Google Scholar]
  • Kimura MK, Kabeya D, Saito T, et al. Effects of genetic and environmental factors on clonal reproduction in old-growth natural populations of Cryptomeria japonica. Forest Ecology and Management. 2013;304:10–19. [Google Scholar]
  • Kondo T, Tsumura Y, Kawahara T, Okamura M. Paternal inheritance of chloroplast and mitochondrial DNA in interspecific hybrids of Chamaecyparis spp. Breeding Science. 1998;48:177–180. [Google Scholar]
  • Langella O. Populations, a population genetic software (version 1.2.30) 1999. Available from http://bioinformatics.org/~tryphon/populations/
  • Machida H, Arai F. Atlas of tephra in and around Japan. Tokyo: University of Tokyo Press (in Japanese); 1992. [Google Scholar]
  • Maiorano L, Cheddadi R, Zimmermann N, et al. Building the niche through time: using 13,000 years of data to predict the effects of climate change on three tree species in Europe. Global Ecology and Biogeography. 2012;22:302–317. [Google Scholar]
  • McCauley D. The use of chloroplast DNA polymorphism in studies of gene flow in plants. Trends in Ecology and Evolution. 1995;10:198–202. [Abstract] [Google Scholar]
  • McDonald GM. Fossil pollen analysis and the reconstruction of plant invasions. Advances in Ecological Research. 1993;24:67–110. [Google Scholar]
  • McLachlan JS, Clark JS. Reconstructing historical ranges with fossil data at continental scales. Forest Ecology and Management. 2004;197:139–147. [Google Scholar]
  • Metz C. Basic principles of ROC analysis. Seminars in Nuclear Medicine. 1978;8:283–298. [Abstract] [Google Scholar]
  • Miyake N, Nakamura J, Yamanaka M, Nakagawa T, Miyake M. Spatial changes in the distribution of Cryptomeria japonica since the last interstade in Shikoku Island, southwestern Japan (Metasequoia: the legacy of Dr. Shigeru Miki: proceedings of the third International Metasequoia Symposium) Japanese Journal of Historical Botany. 2011;19:61–68. [Google Scholar]
  • Moriguchi Y, Matsumoto A, Saito M, Tsumura Y, Taira H. DNA analysis of clonal structure of an old growth, isolated forest of Cryptomeria japonica in a snowy region. Canadian Journal of Forest Research-Revue Canadienne de Recherche Forestière. 2001;31:377–383. [Google Scholar]
  • Moriguchi Y, Iwata H, Ujino-Ihara T, Yoshimura K, Taira H, Tsumura Y. Development and characterization of microsatellite markers for Cryptomeria japonica D.Don. Theoretical and Applied Genetics. 2003;106:751–758. [Abstract] [Google Scholar]
  • Murai S. Kokudo Saiken Zourin Gijutsu Kouenshu. 1947. Major forestry tree species in the Tohoku region and their varietal problems Aomori-rinyukai, 131–151 (in Japanese)
  • Nei M. Molecular evolutionary genetics. New York: Columbia University Press; 1987. [Google Scholar]
  • Ohba K, Iwakawa M, Okada Y, Murai M. Paternal transmission of a plastid anomaly in some reciprocal crosses of sugi, Cryptomeria japonica D. Don. Silvae Genetica. 1971;20:101–107. [Google Scholar]
  • Ohnishi N, Uno R, Ishibashi Y, Tamate H, Oi T. The influence of climatic oscillations during the Quaternary Era on the genetic structure of Asian black bears in Japan. Heredity. 2009;102:579–589. [Abstract] [Google Scholar]
  • Opgenoorth L, Vendramin GG, Mao K, et al. Tree endurance on the Tibetan Plateau marks the world's highest known tree line of the Last Glacial Maximum. New Phytologist. 2010;185:332–342. [Abstract] [Google Scholar]
  • Parducci L, Jørgensen T, Tollefsrud MM, et al. Glacial survival of boreal trees in Northern Scandinavia. Science. 2012;335:1083–1086. [Abstract] [Google Scholar]
  • Peakall R, Smouse PE. GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Notes. 2006;6:288–295. [Europe PMC free article] [Abstract] [Google Scholar]
  • Pearson R, Dawson T. Predicting the impacts of climate change on the distribution of species: are bioclimate envelope models useful? Global Ecology and Biogeography. 2003;12:361–371. [Google Scholar]
  • Petit RJ, Hampe A. Some evolutionary consequences of being a tree. Annual Review of Ecology, Evolution, and Systematics. 2006;37:187–214. [Google Scholar]
  • Petit RJ, Aguinagalde I, de Beaulieu JL, et al. Glacial refugia: hotspots but not melting pots of genetic diversity. Science. 2003;300:1563–1565. [Abstract] [Google Scholar]
  • Piry S, Luikart G, Cornuet JM. Computer note. BOTTLENECK: a computer program for detecting recent reductions in the effective size using allele frequency data. Journal of Heredity. 1999;90:502–503. [Google Scholar]
  • Pornon A, Escaravage N, Till-Bottraud I, Doche B. Variation of reproductive traits in Rhododendron ferrugineum L. (Ericaceae) populations along a successional gradient. Plant Ecology. 1997;130:1–11. [Google Scholar]
  • Pornon A, Escaravage N, Thomas P, Taberlet P. Dynamics of genotypic structure in clonal Rhododendron ferrugineum (Ericaceae) populations. Molecular Ecology. 2000;9:1099–1111. [Abstract] [Google Scholar]
  • Pritchard JK, Stephens M, Donnelly P. Inference of population structure using multilocus genotype data. Genetics. 2000;155:945–959. [Europe PMC free article] [Abstract] [Google Scholar]
  • Provan J, Bennett K. Phylogeographic insights into cryptic glacial refugia. Trends in Ecology and Evolution. 2008;23:564–571. [Abstract] [Google Scholar]
  • Qi XS, Chen C, Comes HP, et al. Molecular data and ecological niche modelling reveal a highly dynamic evolutionary history of the East Asian Tertiary relict Cercidiphyllum (Cercidiphyllaceae) New Phytologist. 2012;196:617–630. [Abstract] [Google Scholar]
  • R Development Core Team. R: a language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing; 2012. [Google Scholar]
  • Sakaguchi S, Sakurai S, Yamasaki M, Isagi Y. How did the exposed seafloor function in postglacial northward range expansion of Kalopanax septemlobus? Evidence from ecological niche modelling. Ecological Research. 2010;25:1183–1195. [Google Scholar]
  • Sakaguchi S, Takeuchi Y, Yamasaki M, Sakurai S, Isagi Y. Lineage admixture during postglacial range expansion is responsible for the increased gene diversity of Kalopanax septemlobus in a recently colonised territory. Heredity. 2011;107:338–348. [Europe PMC free article] [Abstract] [Google Scholar]
  • Savolainen O, Bokma F, García-Gil R, Komulainen P, Repo T. Genetic variation in cessation of growth and frost hardiness and consequences for adaptation of Pinus sylvestris to climatic changes. Forest Ecology and Management. 2004;197:79–89. [Google Scholar]
  • Shimizu Y, Ando M, Sakai F. Clonal structure of natural populations of Cryptomeria japonica growing at different positions on slopes, detected using RAPD markers. Biochemical Systematics and Ecology. 2002;30:733–748. [Google Scholar]
  • Stewart JR, Lister AM. Cryptic northern refugia and the origins of the modern biota. Trends in Ecology and Evolution. 2001;16:608–613. [Google Scholar]
  • Stuessy TF, Crawford DJ, Marticorena C. Patterns of phylogeny in the endemic vascular flora of the Juan Fernandez Islands, Chile. Systematic Botany. 1990;15:338–346. [Google Scholar]
  • Stuessy TF, Jakubowsky G, Gómez RS, et al. Anagenetic evolution in island plants. Journal of Biogeography. 2006;33:1259–1265. [Google Scholar]
  • Taira H, Tsumura Y, Tomaru N, Ohba K. Regeneration system and genetic diversity of Cryptomeria japonica growing at different altitudes. Canadian Journal of Forest Research-Revue Canadienne de Recherche Forestière. 1997;27:447–452. [Google Scholar]
  • Takahara H. Distribution history of Cryptomeria forest. In: Yasuda Y, Miyoushi N, editors. Nippon Rettou Syokusei-si. Tokyo: Asakura-Shoten; 1998. pp. 207–223. (in Japanese) [Google Scholar]
  • Takahashi T, Tani N, Taira H, Tsumura Y. Microsatellite markers reveal high allelic variation in natural populations of Cryptomeria japonica near refugial areas of the last glacial period. Journal of Plant Research. 2005;118:83–90. [Abstract] [Google Scholar]
  • Takahashi T, Tani N, Niiyama K, Yoshida S, Taira H, Tsumura T. Genetic succession and spatial genetic structure in a natural old growth Cryptomeria japonica forest revealed by nuclear and chloroplast microsatellite markers. Forest Ecology and Management. 2008;255:2820–2828. [Google Scholar]
  • Takayama K, Sun B-Y, Stuessy TF. Genetic consequences of anagenetic speciation in Acer okamotoanum (Sapindaceae) on Ullung Island, Korea. Annals of Botany. 2012;109:321–330. [Europe PMC free article] [Abstract] [Google Scholar]
  • Tanaka N, Nakao K, Tsuyama I, Higa M, Nakazono E, Matsui T. Predicting the impact of climate change on potential habitats of fir (Abies) species in Japan and on the East Asian continent. Procedia Environmental Sciences. 2012;13:455–466. [Google Scholar]
  • Tani N, Takahashi T, Ujino-Ihara T, Iwata H, Yoshimura K, Tsumura Y. Development and characteristics of microsatellite markers for sugi (Cryptomeria japonica D. Don) derived from microsatellite-enriched libraries. Annals of Forest Science. 2004;61:569–575. [Google Scholar]
  • Thuiller W, Araujo M, Lavorel S. Generalized models vs. classification tree analysis: predicting spatial distributions of plant species at different scales. Journal of Vegetation Science. 2003;14:669–680. [Google Scholar]
  • Toda R, Sato T. Variety, breeding and the expression of the characteristics of the variety of Cryptomeria japonica. In: Sakaguchi M, editor. Sugi no subete. Tokyo: Zenkoku Ringyo Fukyu Kyokai; 1969. pp. 64–70. (in Japanese) [Google Scholar]
  • Tsuda Y, Ide Y. Wide-range analysis of genetic structure of Betula maximowicziana, a long-lived pioneer tree species and noble hardwood in the cool temperate zone of Japan. Molecular Ecology. 2005;14:3929–3941. [Abstract] [Google Scholar]
  • Tsukada M. The history of Cryptomeria japonica (sugi): the last 15000 years. Kagaku. 1980;50:538–546. (in Japanese) [Google Scholar]
  • Tsukada M. Cryptomeria japonica: glacial refugia and late-glacial and postglacial migration. Ecology. 1982a;63:1091–1105. [Google Scholar]
  • Tsukada M. Late-Quaternary development of the Fagus forest in the Japanease archipelago. Japanease Journal of Ecology. 1982b;32:113–118. [Google Scholar]
  • Tsumura Y, Yoshimura K, Tomaru N, Ohba K. Molecular phylogeny of conifers using RFLP analysis of PCR-amplified specific chloroplast genes. Theoretical and Applied Genetics. 1995;91:1222–1236. [Abstract] [Google Scholar]
  • Tsumura Y, Kado T, Takahashi T, Tani N, Ujino-Ihara T, Iwata H. Genome scan to detect genetic structure and adaptive genes of natural populations of Cryptomeria japonica. Genetics. 2007;176:2393–2403. [Europe PMC free article] [Abstract] [Google Scholar]
  • Tsumura Y, Uchiyama K, Moriguchi Y, Ueno S, Ihara-Ujino T. Genome scanning for detecting adaptive genes along environmental gradients inthe Japanese conifer, Cryptomeria japonica. Heredity. 2012;109:349–360. [Europe PMC free article] [Abstract] [Google Scholar]
  • Tsuyama I, Nakao K, Higa M, Matsui T, Tanaka N. What controls the global distribution of Japanese endemic hemlock, Tsuga diversifolia? Footprint of climate in the glacial periods on the current habitat occupancy. Journal of Forest Research. 2014;19:154–165. [Google Scholar]
  • Ueno S, Moriguchi Y, Uchiyama K, et al. A second generation framework for the analysis of microsatellites in expressed sequence tags and the development of EST-SSR markers for a conifer, Cryptomeria japonica. BMC Genomics. 2012;13:136. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wakeley J. Coalescent theory: an introduction. Greenwood Village: Roberts & Company Publishers; 2008. [Google Scholar]
  • Walter R, Epperson B. Geographic pattern of genetic variation in Pinus resinosa: area of greatest diversity is not the origin of postglacial populations. Molecular Ecology. 2001;10:103–111. [Abstract] [Google Scholar]
  • Weir BS, Cockerham CC. Estimating F-statistics for the analysis of population structure. Evolution. 1984;38:1358–1370. [Abstract] [Google Scholar]
  • Whittaker RJ, Triantis KA, Ladle RJ. A general dynamic theory of oceanic island biogeography. Journal of Biogeography. 2008;35:977–994. [Google Scholar]
  • Widmer A, Lexer C. Glacial refugia: sanctuaries for allelic richness, but not for gene diversity. Trends in Ecology and Evolution. 2001;16:267–269. [Abstract] [Google Scholar]
  • Worth J, Sakaguchi S, Tanaka N, Yamasaki M, Isagi Y. Northern richness and southern poverty: contrasting genetic footprints of glacial refugia in the relictual tree Sciadopitys verticillata (Coniferales: Sciadopityaceae) Biological Journal of the Linnean Society. 2013;108:263–277. [Google Scholar]
  • Worth J, Yokogawa M, Pérez-Figueroa A, et al. Conflict in outcomes for conservation based on population genetic diversity and genetic divergence approaches: a case study in the Japanese relictual conifer Sciadopitys verticillata (Sciadopityaceae) 2014. Conservation Genetics (in press)
  • Yahara T, Ohba H, Murata J, Iwatsuki K. Taxonomic review of vascular plants endemic to Yakushima Island. Journal of the Faculty of Science, the University of Tokyo Selection III. 1987;14:69–119. [Google Scholar]
  • Yamanaka M, Ishikawa S, Sugawara K. Palynological studies of quaternary sediments in northeast Japan. VII. Shiriya-zaki moor in Shimokita Peninsula. Ecological Review. 1990;22:33–47. [Google Scholar]
  • Yasue M, Ogiyama K, Suto S, Tsukahara H, Miyahara F, Ohba K. Geographical differentiation of natural Cryptomeria stands analyzed by diterpene hydrocarbon constituents of individual trees. Journal of the Japanese Forestry Society. 1987;69:152–156. [Google Scholar]
  • Yoshida A, Takeuti S. Quantitative reconstruction of palaeoclimate from pollen profiles in northeastern Japan and the timing of a cold reversal event during the Last Termination. Journal of Quaternary Science. 2009;24:1006–1015. [Google Scholar]
  • Yun JH, Nakao K, Tsuyama I, et al. Does future climate change facilitate expansion of evergreen broad-leaved tree species in human-disturbed landscape of the Korean Peninsula? Implication for monitoring design of the impact assessment. Journal of Forest Research. 2014;19:174–183. [Google Scholar]
  • Zou J, Sun Y, Li L, et al. Population genetic evidence for speciation pattern and gene flow between Picea wilsonii, P. morrisonicola and P. neoveitchii. Annals of Botany. 2013;112:1829–1844. [Europe PMC free article] [Abstract] [Google Scholar]

Articles from Annals of Botany are provided here courtesy of Oxford University Press

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/2831624
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/2831624

Article citations


Go to all (14) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.