Wind-dispersed seeds blur phylogeographic breaks: The complex evolutionary history of Populus lasiocarpa around the Sichuan Basin
Xue Lia,1, Markus Ruhsamb,1, Yi Wanga, Hong-Ying Zhanga, Xiao-Yan Fana, Lei Zhanga, Jing Wanga,**, Kang-Shan Maoa,*     
a. Key Laboratory of Bio-Resource and Eco-Environment of Ministry of Education, College of Life Sciences, State Key Laboratory of Hydraulics and Mountain River Engineering, Sichuan University, Chengdu 610065, Sichuan, PR China;
b. Royal Botanic Garden Edinburgh, 20A Inverleith Row, Edinburgh EH3 5LR, UK
Abstract: The strength of phylogeographic breaks can vary among species in the same area despite being subject to the same geological and climate history due to differences in biological traits. Several important phylogeographic breaks exist around the Sichuan Basin in Southwest China but few studies have focused on wind-dispersed plants. Here, we investigated the phylogeographic patterns and the evolutionary history of Populus lasiocarpa, a wind-pollinated and wind-dispersed tree species with a circum-Sichuan Basin distribution in southwest China. We sequenced and analyzed three plastid DNA fragments (ptDNA) and eight nuclear microsatellites (nSSRs) of 265 individuals of P. lasiocarpa from 21 populations spanning the entire distribution range. Distribution patterns based on nSSR data revealed that there are three genetic groups in P. lasiocarpa. This is consistent with the three phylogeographic breaks (Sichuan Basin, the Kaiyong Line and the 105°E line), where the Sichuan basin acts as the main barrier to gene flow between western and eastern groups. However, the distribution pattern based on ptDNA haplotypes poorly matched the phylogeographic breaks, and wind-dispersed seeds may be one of the main contributing factors. Species distribution modelling suggested a larger potential distribution in the last glacial maximum with a severe bottleneck during the last interglacial. A DIYABC model also suggested a population contraction and expansion for both western and eastern lineages. These results indicate that biological traits are likely to affect the evolutionary history of plants, and that nuclear molecular markers, which experience higher levels of gene flow, might be better indicators of phylogeographic breaks.
Keywords: Phylogeography    Sichuan basin    Populus lasiocarpa    Kaiyong line    Demographic history    
1. Introduction

The mountains of Southwest China are one of the world's biodiversity hotspots (Myers et al., 2000). Understanding the phylogeographic patterns within species, and the evolutionary forces that produced them, is a major objective of evolutionary biologists (Crisp et al., 2011; Usinowicz et al., 2017; Mao et al., 2021). The complex topography and environmental gradients generated by the uplift of the Qinghai-Tibet Plateau (QTP) in Southwest China can restrict dispersal and lead to diversification and intraspecific differentiation of mountain plants. Quaternary climate oscillations, however, may have either facilitated intraspecific differentiation by strengthening existing geographic barriers or may have enhanced gene flow between differentiated populations by bringing them into secondary contact. Recent studies have suggested that four general phylogeographic breaks exist in this region, including the Mekong-Salween Divide, the Tanaka-Kaiyong Line, the Sichuan Basin, and the ca. 105°E line (Ye et al., 2017). These may have existed as long-time gene flow barriers between intraspecific lineages, resulting in their differentiation. However, the geographic location of phylogeographic breaks is not always consistent with known geographic barriers (Soltis et al., 1997; Bond et al., 2001; Puorto et al., 2001; Song et al., 2021), with species-specific biological characteristics likely to be the main confounding factor (Irwin et al., 2001). Therefore, external (e.g., geological and climate history) and intrinsic factors (e.g., biological traits) may work simultaneously and interact with each other to produce present-day distribution patterns of intraspecific genetic variation (Hewitt, 2000; Macqueen et al., 2011).

The Sichuan Basin, which was formed during the middle Miocene (~13 million years ago (Mya)), is one of the most intriguing geographic features of Southwest China (Liu and Xu, 2003). The Basin is a lowland plain surrounded by mountains – the Hengduan Mountains in the west, the Qinling Mountains in the north, the Daba Mountains in the northeast, the Wushan Mountains in the east, the Dalou Mountains in the southeast, and the Yunnan-Guizhou Plateau in the southwest. The complex topographical features of these mountains have created diverse habitats (He and Jiang, 2014) along steep ecological gradients (Liu et al., 2014; Favre et al., 2015) that harbor great species diversity and endemism. Additionally, they have provided potential refugia for plants during climatic extremes, such as the Quaternary glacial cycles (Wen et al., 2017). However, the great elevational differences cause considerable habitat heterogeneity between the basin and the surrounding mountains, resulting in a "soft" genetic barrier for the dispersal of some species and impeding gene flow between populations (Bennett and Provan, 2008; He et al., 2019).

Many previous studies have investigated the phylogeographic patterns of species distributed around the Sichuan Basin (Qiu et al., 2009; Qi et al., 2012; Xie et al., 2012; He and Jiang, 2014; Sun et al., 2014; Ma et al., 2015; Qiao et al., 2018; Wang et al., 2021). For example, phylogeographic analyses have suggested that populations of Davidia involucrata were divided into an eastern and western lineage by the Sichuan Basin, which also separates the Sino-Himalayan and the Sino-Japanese Forest subkingdoms (Ma et al., 2015). Similar phylogeographic breaks have also been identified in Primula ovalifolia (Xie et al., 2012) and Dysosma versipellis (Qiu et al., 2009; Guan et al., 2010). The Sichuan Basin also acts as geographic barrier for closely related species pairs, such as Dipelta floribunda and D. yunnanensis (Tian et al., 2020). Because the genetic differentiation of these species occurred significantly later than the formation of the Sichuan Basin, the phylogeographic structure of these plants may be related to the diverse topography of the Sichuan Basin as well as a changing climate, rather than to the formation of the Sichuan Basin. Furthermore, a few taxa with a ring-shaped distribution around the Sichuan Basin have been studied, such as Tetracentron sinense (Sun et al., 2014), Odorrana margaratae (Qiao et al., 2018), Apodemus draco (Wang et al., 2021), Phoebe zhennan (Xiao et al., 2020) and Cercidiphyllum japonicum (Qi et al., 2012).

In addition to the phylogeographic break of the Sichuan Basin, there are two other phylogeographic breaks, the Kaiyong Line and the 105°E line, which traverse the mountains around the Sichuan Basin (Ye et al., 2017). The "Tanaka-Kaiyong Line" (TKL) is a major phytogeographic boundary in Southwest China composed of the Tanaka line (Tanaka, 1954), located in Yunnan Province, and the Kaiyong line (Lang, 1994), located in Sichuan Province (Li and Li, 1997). It was hypothesized that the TKL is a result of the uplift of QTP and the action of different monsoon systems on areas located to the west and east of the line (Li and Li, 1997). Uplift in the Hengduan Mountains region, located on the southeastern margin of the QTP, is generally believed to have been recent, occurring mainly between the late Miocene and late Pliocene (Zhou et al., 2006; Royden et al., 2008; Wang et al., 2012, 2014; Meng et al., 2014; Deng and Ding, 2015; Fang et al., 2020; Mao et al., 2021). The Asia monsoon system, triggered by the uplift of QTP, led to dramatic climate change in China (Shi et al., 1999; Spicer et al., 2020), and the current monsoon regimes, which differ, were established on either side of the TKL after or from the Late Pleistocene (Wang, 1994; Mitsui et al., 2008; Fan et al., 2013). Strong genetic differentiation across the TKL has been reported previously (Fan et al., 2013; Wang et al., 2014; Tian et al., 2015; Zhao and Gong, 2015; Zhao and Zhang, 2015; Cheng et al., 2021). Some researchers have proposed that the TKL is a monsoon-driven barrier to present-day plant dispersal (Fan et al., 2013; Tian et al., 2015; Zhao and Zhang, 2015). However, in Cephalotaxus oliveri (Wang et al., 2014) and Apodemus draco (Wang et al., 2021), the intraspecific divergence of species was found to be related to the uplift of Yunnan-Guizhou Plateau and QTP.

The 105°E line coincides with the boundary of two floristic subkingdoms, the Sino-Japanese and the Sino-Himalayan Forest subkingdoms (Wu and Wu, 1996; Ye et al., 2017). The Sino-Japanese Forest subkingdom is located in the eastern, warm-to-cold temperate hilly and low-plain areas (0–500 m a.s.l.), a region presently under a Pacific monsoon climate (Qiu et al., 2011). The Sino-Himalayan Forest subkingdom, which has both Pacific and Indian monsoon regimes, consists of a mosaic of plateaus, mountains, basins, river valleys, and deep gorges with wide ranging elevations (500–4000 m a.s.l.) (Wu and Wu, 1996; Qiu et al., 2011). Many phylogeographic studies of plants have shown that a significant genetic barrier exists from 105 to 108°E (Yan et al., 2012; Guo et al., 2014; Sun et al., 2014; Xu et al., 2015). Despite their documented role as barriers to gene flow, whether these three phylogeographic breaks (i.e., the Sichuan Basin, the Kaiyong Line, and the 105°E line) produce common effects on different plant species and molecular markers from different genomes (nuclear, plastid and mitochondrial genomes) of the same species, and why, await further investigation.

Populus lasiocarpa Oliv. is a wind-pollinated and wind-dispersed deciduous tree species that occurs in temperate forests between 1700 and 3500 m around the Sichuan Basin (Luan et al., 2011; Tang et al., 2013). Its distribution, which is unique for poplar germplasm resources in southwest China, extends to Hubei, Shanxi, Guizhou, and Yunnan provinces (Jiang et al., 2015). The species has the largest leaf area in the genus Populus L., and possesses well-developed perianths (Hong et al., 1987), which make it an excellent candidate for exploring the relative contributions of extrinsic (e.g., geological and climate history) versus intrinsic (e.g., biological traits) factors in shaping distribution patterns of genetic variation within species.

In this study, we asked the following questions: (a) Does the distribution pattern of genetic variation in Populus lasiocarpa reflect known phylogeographic breaks around the Sichuan Basin? (b) How have the evolutionary history during the Quaternary and the biological traits of P. lasiocarpa affected its phylogeographic pattern? To answer these questions, we used three plastid fragments (ptDNA) and eight nuclear microsatellite loci markers (nSSRs) to examine the genetic structure of 265 individuals of P. lasiocarpa from 21 populations.

2. Materials and methods 2.1. Plant sampling collection

A total of 265 individuals from 21 populations between 816 and 3000 m a.s.l. was sampled, covering most of the distributional range of Populus lasiocarpa around the Sichuan Basin (Fig. 1; Table S1). Between 6 and 17 trees were sampled from each population, with all sampled individuals at least 100 m apart from one another. Fresh leaves were collected in silica gel and voucher specimens of each population were deposited at the Sichuan University Herbarium (SZ).

Fig. 1 A map showing the sampling locations of 21 populations of Populus lasiocarpa (green discs).
2.2. DNA extraction, amplification, and sequencing

Total genomic DNA was extracted using the CTAB method (Doyle and Doyle, 1987). We successfully sequenced three plastid markers that show high levels of intraspecific variability (matK, trnG-psbK and psbK-psbI; Table S2; Schroeder et al., 2012) in 193 individuals. A set of eight nuclear microsatellite loci were used to genotype 265 individuals (Table S2). Polymerase chain reaction (PCR) for all ptDNA fragments and nSSR loci was performed in 25 μL volumes, containing 50–100 ng DNA, 0.5 mM of each dNTP, 0.5 μL of each primer, and 1 × Taq buffer and 0.5 units of Taq polymerase (Beijing ComWin Biotech Co., Ltd., Chengdu, China). Reactions were run on a Master Cycler Pro thermal cycler (Eppendorf, Hamburg, Germany) using the following protocol: one cycle at 94 ℃ for 5 min, followed by 34 cycles at 95 ℃ for 30 s, 56 ℃ for 30 s, and 72 ℃ for 50 s, with a final extension at 72 ℃ for 5 min. The PCR products were checked on 1% agarose gels and then sent to Beijing Genomics Institute (Chengdu, China) for DNA sequencing and nSSR genotyping.

2.3. nSSR data analyses

Microsatellite data were scored in GeneMarker (Softgenetics, Pennsylvania, USA) and then corrected by the FlexiBin Excel macro (Amos et al., 2007). Allele size was scored at each locus and checked for possible genotyping errors, allelic dropout, and null alleles using CERVUS v.3.0 (Kalnowskl et al., 2007). Two loci (peussr-56336, peussr-83115) with high-frequency null alleles (F [Null] > 0.4) were discarded, the remaining eight nSSR loci (Table S2) were used to estimate genetic diversity indices employing GenAlEx v.6.5 (Peakall and Smouse, 2012). The Bayesian clustering method implemented in STRUCTURE v.2.3.4 was used to investigate population structure in Populus lasiocarpa (Pritchard et al., 2000) with the following parameters: K from 1 to 10 with twenty replicates each, and 1, 500, 000 MCMC cycles after 500, 000 burn-in cycles. The optimal K value was determined using two methods, the likelihood estimate (Pritchard et al., 2000) and the ΔK statistic method (Evanno et al., 2005). STRUCTURE results were summarized and visualized using Structure Harvester (Earl and vonHoldt, 2012). To cross-validate the results of STRUCTURE, we also conducted a Principal Coordinates Analysis (PCoA) using GenAlEx v.6.5 (Peakall and Smouse, 2012). In addition, to assess the spatial population structure, we conducted a spatial analysis of molecular variance using SAMOVA v.2.0 (Dupanloup et al., 2002). We ran the SAMOVA program 100 times simulating the annealing process for K groups (K = 2–10).

2.4. ptDNA sequence analyses

Subsequently, ptDNA sequences were edited, aligned, manually checked, and concatenated with ClustalW in MEGA v.5.0 (Tamura et al., 2011). Insertions/deletions (indels, excluding mononucleotide repeats) were encoded by the software Gapcoder (Young and Healy, 2003), and the 0/1 characters (absence/presence, except '–' gaps after coding) were replaced manually by A/T to use indel information (Havrdová et al., 2015). Haplotype variant sites were detected using DNASP v.5.10 (Librado and Rozas, 2009). NETWORK v.4.6 was used to infer network relationships between ptDNA haplotypes based on sequence variation (Bandelt et al., 1999). In addition, to assess the spatial population structure, we conducted a spatial analysis of molecular variance using SAMOVA v.2.0 (Dupanloup et al., 2002). We ran the SAMOVA program 100 times simulating the annealing process for K groups (K = 2–10). The optimal number of groups (k) was determined based on the highest variance between groups (FCT), incorporating information on haplotype divergence and geographical proximity. Finally, to investigate distribution patterns of genetic variation within the ptDNA data set, analyses of molecular variance (AMOVA) were carried out in ARLEQUIN v.3.5 (Excoffier and Lischer, 2010), with significance tests based on 1000 permutations. Genetic variation was hierarchically partitioned into three levels: among groups, among populations within group, and within populations.

To calculate divergence times, we adopted a two-step approach in BEAST v.1.7.5 (Drummond and Rambaut, 2007), using node ages from Zhang et al. (2018) and four species as outgroups: Populus euphratica, P. wilsonii, P. davidiana, and P. adenopoda (Zhang et al., 2018).

2.5. Genetic barrier analyses

Monmonier's maximum-difference algorithm in BARRIER v.2.2 (Manni et al., 2004) was used to identify biogeographic boundaries or areas exhibiting the largest genetic discontinuities between population pairs. We used nSSR and ptDNA data for genetic barrier analysis. The robustness of these boundaries was assessed by running BARRIER on 100 replicates of population average pairwise difference matrices. These matrices were generated by bootstrapping of haplotype sequences in SEQBOOT (Felsenstein, 2005) and subsequent analyses in ARLEQUIN v.3.5 (Excoffier and Lischer, 2010). One thousand Nei's genetic distance matrices were used to estimate the robustness of gene flow barriers. Nei's genetic distance matrices were produced by bootstrapping loci with microsatellite analyser v.4.05 (Dieringer and Schlötterer, 2003). The genetic groups were distinguished based on genetic discontinuities and were then compared with genetic clades/clusters identified by the phylogenetic tree and STRUCTURE analyses.

2.6. Geographical and environmental effects on genetic divergence

Multiple processes such as isolation-by-environment (IBE) and isolation-by-distance (IBD) may affect genetic variation of populations (Sexton et al., 2014). IBD investigates the correlation between spatial and genetic distance of populations (De Matthaeis et al., 2000; Wang, 2013; Pelletier and Carstens, 2018). IBE implies that genetic divergence increases with environmental heterogeneity (Wang and Bradburd, 2014). Here, three models were examined (Table 1): (a) IBDe, where the geographical distance between populations was represented by a Euclidian distance assuming that Populus lasiocarpa disperses across the basin (coordinates of sample sites were used to calculate "as the crow flies" distances with the pointDistance function in the raster R package; Hijmans and van Etten, 2012); (b) IBDr, the geographical distance between populations was represented by a ring distance (calculated using the ArcMap v.10.2.1), assuming no gene flow across the unoccupied areas across the Sichuan basin; (c) IBE, environmental difference was calculated based on Euclidean distances of the SDM environmental factors (see 2.8. Species distribution modelling (SDM) for details) at each population's locality.

Table 1 Contribution of geographical (Euclidean or ring distances) and environmental variables to the genetic variation of Populus lasiocarpa evaluated by Mantel tests.
Models Mantel test
r p
IBDe 0.462 0.001
IBDr 0.635 0.001
IBE 0.554 0.001
Abbreviations: IBDe, isolation-by-distance, where the geographical distance between populations was represented by a Euclidian distance; IBDr, isolation-by-distance, where the geographical distance between populations was represented by a ring distance; IBE, isolation-by-environment.

We used the philentropy R package (Drost, 2018) to calculate the Euclidean distance between the different environments. Pairwise FST values between populations were calculated in ARLEQUIN v.3.5 (Excoffier and Lischer, 2010). Finally, Spearman's correlation between matrices of genetic, geographical, and environmental distance was assessed by a Mantel test using the Mantel function of the R package vegan (Oksanen et al., 2020) with 999 permutations.

2.7. Demographic history analyses

Mismatch distribution analysis examines the distribution of the observed number of differences between pairs of haplotypes. This distribution is usually multimodal in samples drawn from populations at demographic equilibrium, reflecting the highly stochastic shape of gene trees, whereas it is usually unimodal in populations having passed through a recent demographic or range expansion (Slatkin, 1991; Rogers and Harpending, 1992) with high levels of migration between neighboring demes (Ray et al., 2003). Mismatch distribution was calculated in ARLEQUIN v.3.5 (Excoffier and Lischer, 2010) using the sum of square deviations (SSD) between the observed and the expected mismatches as a test statistic. Arlequin assesses the probability of expansion using the two parameters SSD (Sum of square deviations) and r (Harpending's raggedness index), where the hypothesis of group expansion cannot be rejected if the statistical tests for these two parameters are not significant (p > 0.05).

We used BOTTLENECK v.1.2.02 (Piry et al., 1999) to detect past bottlenecks in three Populus lasiocarpa populations (the eastern, southern and western populations; Table 2) choosing a two-phase model (Di Rienzo et al., 1994) and a two-tailed Wilcoxon signed-rank test to test for significance.

Table 2 Bottleneck analysis for three groups using Wilcoxon test under two phase model.
Geographical population Sample size Expected number of loci with heterozygosity excess Wilcoxon test
Western cluster 71 7.5 0.003906
Southern cluster 55 5 0.015625
Eastern cluster 139 7.75 0.007813

We used the coalescent-based approximate Bayesian computation (ABC) implemented in DIY-ABC v.2.0 (Cornuet et al., 2014) to infer the demographic histories of Populus lasiocarpa. To alleviate the impact of inter-lineage gene flow on the testing of demographic history scenarios, we excluded all individuals that were potential hybrids (according to Fig. 3c, K = 2) from the southern populations and analyzed the demographic history of the eastern and western populations, respectively. Because mismatch distribution analysis and bottleneck analysis indicated a significant expansion and bottleneck in P. lasiocarpa (Fig. S3; Table 2), we tested six models of population size changes (Fig. 2): an old shrinkage followed by expansion (scenario 1: "shrinkage-expansion"); an old expansion followed by shrinkage (scenario 2: "expansion-shrinkage"); a recent expansion followed by shrinkage (scenario 3: "expansion-shrinkage"); expansion followed by shrinkage and a new expansion event (scenario 4: "expansion-shrinkage-expansion"); expansion followed by shrinkage with a small population size and a new expansion event (scenario 5: "expansion-shrinkage-expansion"); a recent shrinkage followed by expansion (scenario 6: "shrinkage-expansion"). We choose the demographic scenarios that best fit the data and estimated the parameters of interest (Table S3). Using a direct approach and logistic regression analyses, the posterior probabilities of all scenarios were calculated and compared. Following Macaya-Sanz et al. (2014) we set the generation time of P. lasiocarpa to 40 years, with a span of 20–60 years.

Fig. 3 A summary of the geographic distribution and genetic clustering between groupings of Populus lasiocarpa based on 8 nuclear microsatellite (nSSR) loci. (a) Population cluster analysis of P. lasiocarpa with plot of the delta K. (b) Population cluster analysis of P. lasiocarpa with plot of the Log probability (L(K)) (mean ± SD). (c) Histogram of the STRUCTURE (K = 2–4) assignment test for 21 populations of P. lasiocarpa based on 8 nSSR loci. (d) Geographic origin of the 21 populations of P. lasiocarpa and their genetic components regarding genetic clusters at the likely K = 3.

Fig. 2 Schematic illustration of the six demographic scenarios (including model parameters) for Populus lasiocarpa tested by DIYABC using the nSSR. The six scenarios assume an old shrinkage followed by expansion (scenario 1: "shrinkage-expansion"); an old expansion followed by shrinkage (scenario 2: "expansion-shrinkage"); a recent expansion followed by shrinkage (scenario 3: "expansion-shrinkage"); expansion followed by shrinkage and a new expansion event (scenario 4: "expansion-shrinkage-expansion"); expansion followed by shrinkage with a small population size and a new expansion event (scenario 5: "expansion-shrinkage-expansion"); a recent shrinkage followed by expansion (scenario 6: "shrinkage-expansion"). Times and effective population size are not strictly to scale.
2.8. Species distribution modelling

Species distribution modelling (SDM) was conducted using the maximum entropy method implemented in MAXENT v.3.2.1 (Phillips et al., 2006) to predict the distribution of potentially suitable habitat for Populus lasiocarpa in five time periods: the Last Glacial Maximum (LGM; ca. 21-thousand years ago (Kya)), the Last Interglacial (LIG; ca. 130 Kya), the Mid-Holocene (MH; ca. 6 Kya), the present, and 2050. Nineteen bioclimatic variables at 2.5 arc-minute resolution were downloaded from the WorldClim database (Fick and Hijmans, 2017). Strong co-linearity between bioclimatic variables may affect the accuracy of the model. Therefore, a Pearson correlation test was performed on these bioclimatic variables across the 21 sampled populations. Seven bioclimatic variables between which all pairwise Pearson correlation coefficients r were > 0.70 (Table S4) were retained and used for subsequent SDM analysis and IBE analyses. The maximum entropy model was simulated for 20 replicates, randomly selecting 20% of the distribution data as the test data, whereas the remaining 80% of the distribution data were used for the training simulation. The maximum number of iterations was set to 5000, and the convergence limit was set to 0.01. The output format was set to "logistic" and the bioclimatic conditions for each grid cell were probabilistic in the range of 0–1. The performance of models was evaluated by checking the area under the Receiving Operating Characteristics (ROC) curve (AUC) value. DIVA-GIS v.7.5 (Hijmans et al., 2001) was used to map the distribution of habitat suitability.

3. Results 3.1. Nuclear genetic differentiation

Using eight nSSR loci, STRUCTURE yielded the highest likelihood for K = 2 (Fig. 3a and b), suggesting the existence of two genetic clusters (Fig. S2a), which occur mainly in the western (populations 1–8) and eastern groups (populations 9–21). For K = 3, there was an eastern group (populations 10 and 12–21), a southern group (populations 7–9 and 11) and a western group (populations 1–6) (Fig. 3c). The PCoA based on individual genetic distances showed a similar population genetic structure, with the southern populations (7, 8, 9, and 11) clustered between the eastern and western populations (Fig. S1b). Spatial genetic analysis indicated that FCT increased to its maximum value of 0.231 at K = 3 (Table S9a). For K = 4 (Figs. 3c and S2b), three populations (2, 3 and 6) in the western group were genetically differentiated to some extent from the other three populations (1, 4 and 5). In the PCoA based on population genetic distance, the distribution of these populations was concordant with geography and generally followed a ring pattern around the barrier (Figs. S1a and S2b).

Averaged over all populations, the mean number of alleles was Aa = 2.73, the mean number of effective alleles was Ae = 1.83, the observed heterozygosity was Ho = 0.25, the expected heterozygosity was He = 0.28, and the inbreeding coefficient at the population level was FIS = 0.14 (Tables S5 and S6). The genetic differentiation index was FST = 0.27 (Table S5).

3.2. Genetic variation of ptDNA sequences

We obtained 1688 bp of concatenated plastid sequences (matK 797 bp; trnG-psbK 476 bp; psbK-psbI 415 bp). A total of 12 haplotypes were detected (Tables S7 and S8), revealing two distinct haplotype clades (Fig. 4c). Clade I comprised all blue haplotypes (H2, H3, H5, H7, H8 and H10). Among these, H2 is the most common haplotype and is present in all but one population (Fig. 4a), indicating that it is the ancestral haplotype of Populus lasiocarpa. Clade II consists of yellow haplotypes (H1 and H4) and red haplotypes (H6, H9, H11 and H12). The yellow haplotypes are mainly represented in the southern populations and the red haplotypes in the western populations (Fig. 4a). Molecular dating suggests that the blue haplotypes diverged from all other clades at ca. 3.66 Ma (Fig. 4c), and that the red diverged from the yellow haplotypes at ca. 2.09 Ma (Fig. 4c). Spatial genetic analysis of ptDNA haplotypes indicated that FCT increased to its maximum value of 0.631 at K = 3 (Table S9b). The grouping pattern of populations corresponding to K = 3 was as follows: 1) population 5; 2) populations 7–9, 11–12 and 17; 3) populations 1–4, 6, 10, 13–16 and 18–21. Based on the SAMOVA groups, AMOVA showed that 75.16% (p < 0.001) of the total variation occurred among groups (Table S10).

Fig. 4 Geographic distribution of haplotype lineages (a) and maximum-parsimony network of 12 haplotypes (b) and the phylogenetic tree of Populus lasiocarpa based on three plastid fragments (c). One haplotype lineage is highlighted in blue and the others in yellow and red. In the network, each haplotype is indicated by an ellipse with their frequencies indicated by the sizes of each ellipse. The colors are concordant with those of the corresponding clades in the phylogenetic tree.
3.3. Genetic barriers

Our BARRIER analysis based on both ptDNA data and nSSR data showed a strong genetic barrier between the southern and western sampling areas (Fig. 5) both received high bootstrap values (100%). In addition, we observed one genetic barrier between populations 2 and 7 (Fig. 5) that had a high support value (100%). There are also other barriers between peripheral populations according to both ptDNA and nSSR data (Fig. 5).

Fig. 5 Map of genetic barriers to (a) plastid haplotypes and (b) nSSR between different sampling areas. In (a) and (b), blue lines show Voronoï tessellation of populations according to geographic locations and red solid lines represent genetic barriers. Circle points represent the locations of the 21 Populus lasiocarpa labeled by the number of their populations.
3.4. Geographic and environmental effects on genetic divergence

The Mantle test based on nSSR data showed that both geographical distance and environmental variables played a considerable role in the differentiation of populations (Table 1). However, the fit was significant when the geographical variable was based on ring (r = 0.635, p = 0.001) rather than Euclidean distances (r = 0.462, p = 0.001), with a significant correlation between genetic variation and ring distance. In addition, compared to the environmental variables (r = 0.554, p = 0.001), the ring distance (r = 0.635, p = 0.001) delivered a stronger effect on the differentiation of populations.

3.5. Demographic history

The mismatch distribution of Populus lasiocarpa was unimodal (Fig. S3), fitting the expected distribution under the sudden expansion model, supporting a significant expansion of P. lasiocarpa (p [SSD] = 0.44, p [RAG] = 0.19). There was a significant heterozygosity excess (Table 2, p < 0.05) under all three models in the western, southern and eastern populations, indicating a putative bottleneck that might have occurred in those population.

In the DIYABC analyses, the scenarios best supported based on direct estimate, logistic regression and PCA plots were "expansion-shrinkage-expansion" (scenario 5) for the eastern populations and "shrinkage-expansion" (scenario 6) for western populations (Figs. S4 and S5). These scenarios were also the least error prone (Table S11). The "expansion-shrinkage-expansion" (scenario 5) best-fit for the eastern populations indicates that Populus lasiocarpa initially expanded before the LGM period but not earlier than the LIG period, ca. 62 Kya (95% CI: 20.2–88.8 Kya), then experienced an obvious reduction in population size (ca. 11.6 Kya), and expanded again before the MH, ca. 6.7 Kya (95% CI: 1.5–11.3 Kya) (Table S12). By contrast, the "shrinkage-expansion" (scenario 6) best-fit for western populations indicates that a strong bottleneck occurred during the postglacial period, ca. 12.0 Kya (95% CI: 5.1–22.6 Kya), and expanded again before the MH, ca. 6.4 Kya (95% CI: 0.47–11.6 Kya) (Table S12).

3.6. Species distribution predicted by modelling

The predictions of potentially suitable distribution areas under the present and historical climate conditions are shown in Fig. 6. All models have high AUC test values (> 0.95), indicating the good performance of all predictions. Variable jackknife analyses suggested that Max Temperature of Warmest Month (56.4%) and Mean Monthly Temperature Range (22.2%) were the environmental variables that contributed the most to the distribution modelling for Populus lasiocarpa (Table S4).

Fig. 6 Ecological niche modelling predicted distributional range for Populus lasiocarpa species at five periods. (a) Photograph of P. lasiocarpa (courtesy of Lei Zhang). Five periods are (b) the last interglacial (LIG), (c) the last glacial maximum (LGM), (d) the middle Holocene (MH), (e) the present time and (f) 2050.

The projected distribution of this species at present encompasses most of the sampling locations. During the LIG period, the range of P. lasiocarpa was narrow and scattered, relative to the current distribution, and it seemed to have been restricted mainly to the west and east of the Sichuan Basin, with a more fragmented distribution in the east (Fig. 6b). The range of P. lasiocarpa expanded from the LIG to the LGM (Fig. 6c) including extended distribution areas northwards and southwards of the Sichuan Basin. There was no obvious expansion or contraction for the distributions of P. lasiocarpa under present (Fig. 6e) and MH (Fig. 6d) climate conditions, yet only a very tiny distribution of the species was predicted for 2050 (Fig. 6f).

4. Discussion

In this study, we used a suite of phylogeographic analytical tools to investigate the evolutionary history of Populus lasiocarpa, a tree species that occurs in the mountains surrounding the Sichuan Basin. Bayesian clustering based on nSSR markers (Fig. 3) indicated that the populations of P. lasiocarpa were divided into three population groups (eastern, southern and western) by three phylogeographic breaks, the Sichuan Basin, the Kaiyong Line, and the 105°E line (Fig. 3d). Distribution patterns based on ptDNA haplotypes poorly matched these phylogeographic breaks (Fig. 4a). A Mantle test suggested that both geographical distance and environmental variables played a considerable role in the differentiation of populations, especially using the ring distance (Table 1). In addition, SDM analyses indicated a much more restricted distribution during the LIG than at present, and a slightly greater area at the LGM than that predicted under current climatic conditions (Fig. 6). Finally, the DIYABC analyses suggested a postglacial range contraction and a middle Holocene expansion for both western and eastern populations (Table S12).

4.1. The influence of multiple forces on the contemporary genetic structure

Results from STRUCTURE analysis using nSSR makers (Fig. 3) indicated that the Sichuan Basin acted as a phylogeographic boundary that divided Populus lasiocarpa populations into two genetic groups, one to the east and another to the west of the basin. In addition, our analyses showed that this geographical barrier led to a ring diversification pattern in P. lasiocarpa. The mountain ranges surrounding the Sichuan Basin in southwestern China are one of 100 top candidate regions for potential ring species (Monahan et al., 2012). Few polytypic taxa are in fact recognized by modern taxonomy as ring species (Irwin et al., 2001). Therefore, "ring diversification" (Monahan et al., 2012) is used to describe the situation in which populations diverge around a geographical barrier in a ring-like manner, regardless of whether reproductive isolation occurs at the terminus, such as for Hyla orientalis (Dufresnes et al., 2016), Odorrana margaratae (Qiao et al., 2018), Tetracentron sinense (Sun et al., 2014), and Apodemus draco (Wang et al., 2021). The populations of P. lasiocarpa are distributed in the mountains surrounding the Sichuan Basin in a ring like fashion (Figs. 3d, 4a, S1, and S2). The results of the SDM clearly show that the Sichuan Basin must have been an unsuitable habitat for P. lasiocarpa in historical and contemporary times, preventing gene flow across the basin (Fig. 6). The genetic divergence of P. lasiocarpa showed a higher correlation with the ring distance than with the Euclidean distance and environmental variables (Table 1), indicating limited dispersal across the barrier. Thus, the Sichuan Basin, which has been a strong geographical barrier since the middle Miocene (~13 Mya) (Liu and Xu, 2003), not only divided P. lasiocarpa into two genetic groups located in the western and eastern parts of the basin but caused the formation of a ring-shaped distribution.

Three groups showed strong isolation signals in the STRUCTURE analyses (K = 3) and PCoA (Fig. S1) based on the nSSR data. The SAMOVA groups based on ptDNA haplotypes and nSSR data also indicated that K = 3 was the best grouping pattern (Table S9). However, the groupings based on ptDNA poorly matched that of nSSRs as the same ptDNA haplotypes, H2 and H1, occur in two or more nSSR groups. The barrier effect of the Sichuan Basin is so strong that pollen and seeds rarely cross it. Hence, in addition to seed dispersal, shared ancestral polymorphism may also explain this pattern. Regardless, nSSR results suggested there are three geographical genetic groups, located at the western, southern, and eastern portions of the basin edge (Figs. 3d and S1a). These groups approximately reflect three subregions: (a) the Wushan and Daba Mountains at the eastern edge of the Sichuan Basin; (b) the Great Lou Mountains and the Great Liangshan Mountains at the southwestern edge of the basin; and (c) the Great Snowy Mountains, Qionglai Mountains and Minshan Mountains, located on the edge of the Hengduan Mountains and the QTP. Moreover, BARRIER analysis showed a strong genetic barrier between populations 2 and 7 based on both nSSR and ptDNA data (Fig. 5), which coincides with the Kaiyong Line. We observed one genetic barrier between populations 11 and 12 that coincided with the 105°E line (Fig. 2d), but it had a low support value (Fig. 5b). This low support value may be explained by the abnormality that occurs when genetic and geographic distances are coupled, as BARRIER analysis is less sensitive to barriers between long-distance populations. Interestingly, the geopositions of three genetic groups in Populus lasiocarpa correspond to three phylogeographical breaks, including the Sichuan Basin, the Kaiyong line and the 105°E line. Therefore, for P. lasiocarpa, the Sichuan Basin is a stronger geographical barrier, and there are also important genetic barriers at the Kaiyong line and the 105°E line.

Molecular dating in our study suggested that the intraspecific divergence time of haplotypes H1, H4 and haplotypes H9, H11, H12 on both sides of the Kaiyong line occurred approximately 2.09 Mya (Fig. 4c). This intraspecific divergence time coincided with uplift events of the QTP, which are known as the Qingzang Movement (~3.6–1.7 Mya) (Li and Fang, 1999). The intraspecific divergence time of A.podemus draco surrounding the Sichuan Basin (1.7 Mya) also coincided with the tectonic events of the QTP (Wang et al., 2021). The spatial and temporal concordance among these species suggests that tectonic events in the QTP were the same drivers in building their genetic differentiation. Thus, our results support the hypothesis that the historical tectonic events may be factors contributing to the formation of genetic barriers at the Kaiyong line. At the same time, we cannot rule out the possibility of monsoon driven hypothesis, because the latitude and the annual precipitation of western populations is higher than those of the southern populations; however, we found that a glacial driven hypothesis is less likely considering the ecological niche model of the LGM (Fig. 6c), where the potential distribution is even larger compared to the Current model (Fig. 6e).

Similar to Populus lasiocarpa, research on phylogeography of plants distributed around the Sichuan Basin has indicated genetic differentiation of some plants at the 105°E line, including Fagus hayatae (Ying et al., 2016), Liriodendron chinense (Yang et al., 2019), and Davidia involucrata (Ma et al., 2015). Notably, SDM analysis predicted low habitat suitability and gaps in the distribution of P. lasiocarpa around the 105°E line during three time periods (LGM, MH, and the present). The genetic differentiation of P. lasiocarpa at the 105°E line might be the result of different climatic and topographical processes on either side (Qiu et al., 2011; Ye et al., 2017).

Interestingly, our nSSRs data suggest the existence of a gene flow barrier at the 105°E line between populations 11 and 12 (Fig. 2d), yet a strong genetic barrier between populations 12 and 13 was supported by the ptDNA data (Fig. 5a). In plants, seed dispersal is often considerably less effective than pollen dispersal (Petit et al., 2005). Therefore, maternally inherited organelle markers (ptDNA or mtDNA) that are only dispersed by seeds should be more frequently introgressed (Petit and Excoffier, 2009). Moreover, due to the prevalence of sex-biased dispersal (Greenwood, 1980), different DNA regions are subject to varying levels of gene flow as a consequence of their mode of inheritance (biparental, maternal or paternal). Based on our nSSR data, we speculate that pollen and seeds of Populus lasiocarpa likely crossed this phylogeographic break at the 105°E line with the help of the monsoons from both the Indian and Pacific oceans. We suggest the idea that nuclear molecular markers, which have higher rates of gene flow, might be better indicators of phylogeographic breaks than organelle markers, which have lower rates of gene flow, in part because higher rates of intraspecific or local gene flow can prevent interspecific or between-lineage introgression (Petit and Excoffier, 2009). This is evident in P. lasiocarpa because wind-dispersed seeds probably facilitate introgression of ptDNA between geographical genetic groups compared to seeds dispersed by gravity, and large effective population size of trees favor the retention of these introgressed haplotypes. In addition to ancestral polymorphism and wind-dispersed seeds, hybridization-induced plastid capture events between closely related species may also have influenced the distribution pattern of ptDNA haplotypes. Wang et al. (2020) identified some gene flow and hybridization between P. lasiocarpa and P. wilsonii. The evolutionary history of P. lasiocarpa and its relative species can be studied in depth in the future using genomic-based approaches to better reveal hybridization and resulting introgression.

4.2. Infrequent demographic histories of Populus lasiocarpa

Our SDM analyses indicated a much more restricted distribution of Populus lasiocarpa during the LIG than at present, and a slightly greater area at the LGM than that predicted under current climatic conditions (Fig. 6). The mismatch distribution supported a significant expansion (Fig. S3), and the bottleneck analysis indicated that P. lasiocarpa experienced a recent genetic bottleneck (Table 2). Consistent with this, the DIYABC analysis suggested that both western and eastern groups of P. lasiocarpa experienced a bottleneck 11, 600 to 120, 000 years ago, which matches the Younger Dryas (Cheng et al., 2020), an abrupt increase of paleotemperature during postglacial periods; a demographic expansion was also detected for both western and eastern groups around the middle Holocene, 6440–6720 years ago, which is consistent with previous observations that paleotemperature started to decline across much of the globe around 5000–7000 years ago (Steig, 1999).

These findings contradict the scenario where most plants experienced expansion during the LIG and then contraction during the LGM in southern China (Qu et al., 2009; Hofmann, 2012; Zhou et al., 2013; Wang et al., 2017, 2018). However, the climatic changes in the Sichuan Basin were not as pronounced as in other regions (such as Europe and North America) (Jablonski, 1993; Kose, 2014), and some species even had a constant distributional range during this period (Chen et al., 2015; Cao et al., 2016; Wang et al., 2021). As refugia, the mountains of Southwest China contributed to a more stable distribution pattern by providing diverse micro-environments that prevented over-heating or over-cooling (Wen et al., 2017). Paleoclimatic data have indicated that during the LIG the temperature was at least 5 ℃ higher than at present (Shi et al., 1999; Li and Pan, 2002; Li and Kang, 2006). For some cold-adapted or moisture-loving plants, climate warming such as that during the LIG or postglacial period could have led to range contraction instead. Taxus wallichiana and Picea likiangensis in the adjacent areas of the QTP have already been affected in this way in that their current predicted distributions have been reported to be smaller than those predicted at the LGM (Li et al., 2013; Liu et al., 2013). Biological characteristics of Populus lasiocarpa indicate its sensitivity to high temperatures: P. lasiocarpa has the largest leaf area in poplars, about 15–30 cm long and 12–20 cm wide (Hong et al., 1987), which makes it intolerant of drought and high temperatures due to strong transpiration and respiration (Rozendaal et al., 2006; Picotte et al., 2007; Li et al., 2011); in addition, variable jackknife analyses suggest that Max Temperature of Warmest Month (56.4%) was the environmental variable that contributed most to potential distribution modelling for P. lasiocarpa (Table S4). Therefore, P. lasiocarpa may have experienced a bottleneck event during the postglacial period, and most likely also during the LIG (Fig. 6b), when the paleoclimate was warmer, in contrast to the more stable and milder environment of the mountains surrounding the Sichuan Basin during the LGM that P. lasiocarpa populations might have favored. The demographic expansion of western and eastern groups of P. lasiocarpa following paleotemperature decline around the middle Holocene also suggests that this species favors mild environments. Hence, considering that climate warming during the LIG had a marked effect on the predicted distribution of P. lasiocarpa, coupled with a predicted range contraction of the distribution of this species in 2050 (Fig. 6f), population decline and range contraction of this species in the future cannot be ruled out. Taken together, our findings indicate that the conservation and management of this remarkable and beautiful poplar is an imperative task.

Acknowledgements

We thank Jia-Liang Li and Wen-Jing Tao for their assistance in sample collection. This project was supported by National Natural Science Foundation of China (grants 31971567 and 31622015) and Fundamental Research Funds for the Central Universities (YJ201936, SCU2020D003, SCU2021D006, SCU2022D003).

Author contributions

K.-S.M. and J.W. designed the study. H.-Y.Z., X.-Y.F, X.L. and L.Z. performed the field work; X.L. and H.-Y.Z. analyzed the data; X.L. and Y.W. wrote the initial draft of the manuscript. M.R. contributed new reagent to the manuscript. K.-S.M, X.L., M.R. and J.W. finalized the manuscript, and all authors contributed to interpretation/discussions of the data and revisions of the manuscript. All authors approved the final manuscript for publication.

Data availability statement

All sequences have been deposited in GenBank (accession nos. OL878409–OL878601 for psbK-psbI, OL878602–OL878794 for trnG-psbK, OL878795–OL878987 for matK). Microsatellite data and ptDNA sequences data have been deposited in the Science Data Bank and are available at http://doi.org/10.57760/sciencedb.j00143.00020.

Declaration of competing interest

The authors have no conflict of interest.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.pld.2022.10.003.

References
Amos, W., Hoffman, J.I., Frodsham, A., et al., 2007. Automated binning of microsatellite alleles: problems and solutions. Mol. Ecol. Notes, 7: 10-14.
Bandelt, H.J., Forster, P., Röhl, A., 1999. Median-joining networks for inferring intraspecific phylogenies. Mol. Biol. Evol., 16: 37-48. DOI:10.1093/oxfordjournals.molbev.a026036
Bennett, K.D., Provan, J., 2008. What do we mean by "refugia". Quat. Sci. Rev., 27: 2449-2455. DOI:10.1016/j.quascirev.2008.08.019
Bond, J.E., Hedin, M.C., Ramirez, M.G., et al., 2001. Deep molecular divergence in the absence of morphological and ecological change in the Californian coastal dune endemic trapdoor spider Aptostichus simus. Mol. Ecol., 10: 899-910. DOI:10.1046/j.1365-294X.2001.01233.x
Cao, Y.N., Comes, H.P., Sakaguchi, S., et al., 2016. Evolution of east Asia's Arcto-tertiary relict Euptelea (Eupteleaceae) shaped by late Neogene vicariance and Quaternary climate change. BMC Evol. Biol., 16: 66. DOI:10.1186/s12862-016-0636-x
Chen, J.M., Zhao, S.Y., Liao, Y.Y., et al., 2015. Chloroplast DNA phylogeographic analysis reveals significant spatial genetic structure of the relictual tree Davidia involucrata (Davidiaceae). Conserv. Genet., 16: 583-593. DOI:10.1007/s10592-014-0683-z
Cheng, H., Zhang, H., Spötl, C., et al., 2020. Timing and structure of the Younger Dryas event and its underlying climate dynamics. Proc. Natl. Acad. Sci. U.S.A., 117: 23408-23417. DOI:10.1073/pnas.2007869117
Cheng, S., Zeng, W., Fan, D., et al., 2021. Subtle East-West phylogeographic break of Asteropyrum (Ranunculaceae) in subtropical China and adjacent areas. Diversity, 13: 627. DOI:10.3390/d13120627
Cornuet, J.M., Pudlo, P., Veyssier, J., et al., 2014. DIYABC v2.0: a software to make approximate Bayesian computation inferences about population history using single nucleotide polymorphism, DNA sequence and microsatellite data. Bioinformatics, 8: 1187-1189. DOI:10.1093/bioinformatics/btt763
Crisp, M.D., Trewick, S.A., Cook, L.G., 2011. Hypothesis testing in biogeography. Trends Ecol. Evol., 26: 66-72. DOI:10.1016/j.tree.2010.11.005
De Matthaeis, E., Davolos, D., Cobolli, M., et al., 2000. Isolation by distance in equilibrium and nonequilibrium populations of four Talitrid species in the Mediterranean sea. Evolution, 54: 1606-1613.
Deng, T., Ding, L., 2015. Paleoaltimetry reconstructions of the Tibetan Plateau: progress and contradictions. Natl. Sci. Rev., 2: 417-437. DOI:10.1093/nsr/nwv062
Di Rienzo, A., Peterson, A.C., Garza, J.C., et al., 1994. Mutational processes of simple-sequence repeat loci in human populations. Proc. Natl. Acad. Sci. U.S.A., 91: 3166-3170. DOI:10.1073/pnas.91.8.3166
Dieringer, D., Schlötterer, C., 2003. Microsatellite analyser (MSA): a platform independent analysis tool for large microsatellite data sets. Mol. Ecol. Notes, 3: 167-169. DOI:10.1046/j.1471-8286.2003.00351.x
Doyle, J.J., Doyle, J.L., 1987. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull., 19: 11-15.
Drost, H.G., 2018. Philentropy: similarity and distance quantification between probability functions. J. Open Source Softw., 3: 765. DOI:10.21105/joss.00765
Drummond, A.J., Rambaut, A., 2007. BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evol. Biol., 7: 214. DOI:10.1186/1471-2148-7-214
Dufresnes, C., Litvinchuk, S.N., Leuenberger, J., et al., 2016. Evolutionary melting pots: a biodiversity hotspot shaped by ring diversifications around the Black Sea in the Eastern tree frog (Hyla orientalis). Mol. Ecol., 25: 4285-4300. DOI:10.1111/mec.13706
Dupanloup, I., Schneider, S., Excoffier, L., 2002. A simulated annealing approach to define the genetic structure of populations. Mol. Ecol., 11: 2571-2581. DOI:10.1046/j.1365-294X.2002.01650.x
Earl, D.A., vonHoldt, B.M., 2012. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conserv. Genet. Resour., 4: 359-361. DOI:10.1007/s12686-011-9548-7
Evanno, G., Regnaut, S., Goudet, J., 2005. Detecting the number of clusters of individuals using the software structure: a simulation study. Mol. Ecol., 14: 2611-2620. DOI:10.1111/j.1365-294X.2005.02553.x
Excoffier, L., Lischer, H., 2010. ARLEQUIN suite ver 3.5: a new series of programs to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour., 10: 564-567. DOI:10.1111/j.1755-0998.2010.02847.x
Fan, D.M., Yue, J.P., Nie, Z.L., et al., 2013. Phylogeography of Sophora davidii (Leguminosae) across the 'Tanaka-Kaiyong line', an important phytogeographic boundary in southwest China. Mol. Ecol., 22: 4270-4288. DOI:10.1111/mec.12388
Fang, X.M., Dupont-Nivet, G., Wang, C.S., et al., 2020. Revised chronology of central Tibet uplift (Lunpola basin). Sci. Adv., 6: eaba7298. DOI:10.1126/sciadv.aba7298
Favre, A., Päckert, M., Pauls, S.U., et al., 2015. The role of the uplift of the Qinghai-Tibetan Plateau for the evolution of Tibetan biotas. Biol. Rev., 90: 236-253. DOI:10.1111/brv.12107
Felsenstein, J., 2005. PHYLIP (phylogeny inference package). R package version 3.6. http://evolution.genetics.washington.edu/phylip.html.
Fick, S.E., Hijmans, R.J., 2017. WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. Int. J. Climatol., 37: 4302-4315. DOI:10.1002/joc.5086
Greenwood, P.J., 1980. Mating systems, philopatry and dispersal in birds and mammals. Anim. Behav., 28: 1140-1162. DOI:10.1016/S0003-3472(80)80103-5
Guan, B.C., Fu, C.X., Qiu, Y.X., et al., 2010. Genetic structure and breeding system of a rare understory herb, Dysosma versipellis (Berberidaceae), from temperate deciduous forests in China. Am. J. Bot., 97: 111-122. DOI:10.3732/ajb.0900160
Guo, X.D., Wang, H.F., Bao, L., et al., 2014. Evolutionary history of a widespread tree species Acer mono in East Asia. Ecol. Evol., 4: 4332-4345. DOI:10.1002/ece3.1278
Havrdová, A., Douda, J., Krak, K., et al., 2015. Higher genetic diversity in recolonized areas than in refugia of Alnus glutinosa triggered by continent-wide lineage admixture. Mol. Ecol., 24: 4759-4777. DOI:10.1111/mec.13348
He, K., Gutiérrez, E.E., Heming, N.M., et al., 2019. Cryptic phylogeographic history sheds light on the generation of species diversity in sky-island mountains. J. Biogeogr., 46: 2232-2247. DOI:10.1111/jbi.13664
He, K., Jiang, X.L., 2014. Sky islands of southwest China. I: an overview of phylogeographic patterns. Chin. Sci. Bull., 59: 585-597. DOI:10.1007/s11434-013-0089-1
Hewitt, G., 2000. The genetic legacy of the Quaternary ice ages. Nature, 405: 907-913. DOI:10.1038/35016000
Hijmans, R., van Etten, J., 2012. Geographic Analysis and Modeling with Raster Data. R package version 3.5-2. https://rspatial.org/raster/.
Hijmans, R.J., Guarino, L., Cruz, M., et al., 2001. Computer tools for spatial analysis of plant genetic resources data: 1. DIVA-GIS. Plant Genet. Resour. Newsl. 127, 15-19. https://diva-gis.org/docs/pgr127_15-19.pdf.
Hofmann, S., 2012. Population genetic structure and geographic differentiation in the hot spring snake Thermophis baileyi (Serpentes, Colubridae): indications for glacial refuges in southern-central Tibet. Mol. Phylogenet. Evol., 63: 396-406. DOI:10.1016/j.ympev.2012.01.014
Hong, T., Ma, Z.L., Chen, J.S., 1987. Floral morphology of Populus lasiocarpa Oliv. and its phylogenetic position in Populus. J. Integr. Plant Biol., 29: 236-241.
Irwin, D.E., Irwin, J.H., Price, T.D., 2001. Ring species as bridges between microevolution and speciation. Genetica, 112–113: 223-243.
Jablonski, N.G., 1993. Quaternary environments and the evolution of primates in East Asia, with notes on two new specimens of fossil Cercopithecidae from China. Folia Primatol. Int. J. Primatol., 60: 118-132. DOI:10.1159/000156681
Jiang, X.L., Jia, C., Dai, L.L., et al., 2015. Research on the growth characteristics of Populus lasiocarpa Oliv. In west Sichuan plateau. J. Sichuan For. Sci. Technol., 36: 13-17.
Kalnowskl, S.T., Taper, M.L., Marshall, T.C., 2007. Revising how the computer program CERVUS accommodates genotyping error increases success in paternity assignment. Mol. Ecol., 16: 1099-1106. DOI:10.1111/j.1365-294X.2007.03089.x
Kose, S., 2014. Glacial Advances in Asia, Europe, and North America. In: Smith, C. (Ed. ), Encyclopedia of Global Archaeology. Springer, New York, NY, pp. 3041-3043.
Lang, K.Y., 1994. Studies on the distribution patterns of some significant genera in orchid flora. Acta Phytotaxon. Sin., 32: 328. DOI:10.1515/bmte.1994.39.s1.328
Li, B.Y., Pan, B.T., 2002. Progress in paleogeographic study of the Tibetan Plateau. Geogr. Res., 21: 61-70.
Li, C.L., Kang, S.C., 2006. Review of studies in climate change over the Tibetan Plateau. Acta Geograph. Sin., 61: 327-335.
Li, J., Fang, X., 1999. Uplift of the Tibetan Plateau and environmental changes. Chin. Sci. Bull., 44: 2117-2124. DOI:10.1007/BF03182692
Li, L., Abbott, R.J., Liu, B.B., et al., 2013. Pliocene intraspecific divergence and Plio-Pleistocene range expansions within Picea likiangensis (Lijiang spruce), a dominant forest tree of the Qinghai-Tibet plateau. Mol. Ecol., 22: 5237-5255. DOI:10.1111/mec.12466
Li, X.W., Li, J., 1997. The Tanaka-Kaiyong Line - an important floristic line for the study of the flora of East Asia. Ann. Mo. Bot. Gard., 84: 888-892. DOI:10.2307/2992033
Li, Y.H., Lu, Q., Wu, B., et al., 2011. A review of leaf morphology plasticity linked to plant response and adaptation characteristics in arid ecosystems. Chin. J. Plant Ecol., 36: 88.
Librado, P., Rozas, J., 2009. DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics, 25: 1451-1452. DOI:10.1093/bioinformatics/btp187
Liu, J., Möller, M., Provan, J., et al., 2013. Geological and ecological factors drive cryptic speciation of yews in a biodiversity hotspot. New Phytol., 199: 1093-1108. DOI:10.1111/nph.12336
Liu, J.Q., Duan, Y.W., Hao, G., et al., 2014. Evolutionary history and underlying adaptation of alpine plants on the Qinghai–Tibet Plateau. J. Syst. Evol., 52: 241-249. DOI:10.1111/jse.12094
Liu, Y.K., Xu, C., 2003. Modeling for the burial and subsidence history of the Sichuan Basin. Chin. J. Geophys., 46: 283-290. DOI:10.1002/cjg2.343
Luan, H.H., Su, X.H., Zhang, B.Y., 2011. Research progress in genetic evaluation of Populus L. germplasm resources. Chin. Bull. Bot., 46: 586-595. DOI:10.3724/SP.J.1259.2011.00586
Ma, Q., Du, Y.J., Chen, N., et al., 2015. Phylogeography of Davidia involucrata (Davidiaceae) inferred from cpDNA haplotypes and nSSR data. Syst. Bot., 40: 796-810. DOI:10.1600/036364415X689267
Macaya-Sanz, D., Heuertz, M., López-de-Heredia, U., et al., 2014. The Atlantic–Mediterranean watershed, river basins and glacial history shape the genetic structure of Iberian poplars. Mol. Ecol., 21: 3593-3609.
Macqueen, P., Goldizen, A.W., Austin, J.J., et al., 2011. Phylogeography of the Pademelons (Marsupialia: Macropodidae: Thylogale) in new Guinea reflects both geological and climatic events during the Plio-Pleistocene. J. Biogeogr., 38: 1732-1747. DOI:10.1111/j.1365-2699.2011.02522.x
Manni, F., Guérard, E., Heyer, E., 2004. Geographic patterns of (Genetic, Morphologic, Linguistic) variation: how barriers can be detected by using Monmonier's Algorithm. Hum. Biol., 76: 173-190. DOI:10.1353/hub.2004.0034
Mao, K.S., Wang, Y., Liu, J.Q., 2021. Evolutionary origin of species diversity on the Qinghai–Tibet Plateau. J. Syst. Evol., 59: 1142-1158. DOI:10.1111/jse.12809
Meng, K., Wang, E., Wang, G., 2014. Uplift of the Emei Shan, western Sichuan basin: implication for eastward propagation of the Tibetan Plateau in early Miocene. J. Asian Earth Sci., 115: 29-39. DOI:10.1007/978-94-007-7329-5_3
Mitsui, Y., Chen, S.T., Zhou, Z.K., et al., 2008. Phylogeny and biogeography of the Genus Ainsliaea (Asteraceae) in the Sino-Japanese region based on nuclear rDNA and plastid DNA sequence data. Ann. Bot., 101: 111-124.
Monahan, W.B., Pereira, R.J., Wake, D.B., 2012. Ring distributions leading to species formation: a global topographic analysis of geographic barriers associated with ring species. BMC Biology, 10: 20. DOI:10.1186/1741-7007-10-20
Myers, N., Mittermeler, R.A., Mittermeler, C.G., et al., 2000. Biodiversity hotspots for conservation priorities. Nature, 403: 853-858. DOI:10.1038/35002501
Oksanen, J., Blanchet, F.G., Friendly, M., et al., 2020. Vegan: community ecology package. R package version 2.5-7. https://CRAN.R-project.org/package=vegan.
Peakall, R., Smouse, P.E., 2012. GenAlEx 6.5: genetic analysis in Excel. Population genetic software for teaching and research-an update. Bioinformatics, 28: 2537-2539. DOI:10.1093/bioinformatics/bts460
Pelletier, T.A., Carstens, B.C., 2018. Geographical range size and latitude predict population genetic structure in a global survey. Biol. Lett., 14: 20170566. DOI:10.1098/rsbl.2017.0566
Petit, R.J., Duminil, J., Fineschi, S., et al., 2005. Comparative organization of chloroplast, mitochondrial and nuclear diversity in plant populations. Mol. Ecol., 14: 689-701.
Petit, R.J., Excoffier, L., 2009. Gene flow and species delimitation. Trends Ecol. Evol., 24: 386-393. DOI:10.1016/j.tree.2009.02.011
Phillips, S.J., Anderson, R.P., Schapire, R.E., 2006. Maximum entropy modeling of species geographic distributions. Ecol. Model., 190: 231-259. DOI:10.1504/IJGENVI.2006.010156
Picotte, J.J., Rosenthal, D.M., Rhode, J.M., et al., 2007. Plastic responses to temporal variation in moisture availability: consequences for water use efficiency and plant performance. Oecologia, 153: 821-832. DOI:10.1007/s00442-007-0794-z
Piry, S., Luikart, G., Cornuet, J.M., 1999. Computer note. BOTTLENECK: a computer program for detecting recent reductions in the effective size using allele frequency data. J. Hered., 90: 502-503. DOI:10.1093/jhered/90.4.502
Pritchard, J.K., Stephens, M., Donnelly, P., 2000. Inference of population structure using multilocus genotype data. Genetics, 155: 945-959. DOI:10.1093/genetics/155.2.945
Puorto, G., Da Graça Salomão, M., Theakston, R.D.G., et al., 2001. Combining mitochondrial DNA sequences and morphological data to infer species boundaries: phylogeography of lanceheaded pitvipers in the Brazilian Atlantic forest, and the status of Bothrops pradoi (Squamata: serpentes: Viperidae). J. Evol. Biol., 14: 527-538. DOI:10.1046/j.1420-9101.2001.00313.x
Qi, X.S., Chen, C., Comes, H.P., et al., 2012. Molecular data and ecological niche modelling reveal a highly dynamic evolutionary history of the East Asian Tertiary relict Cercidiphyllum (Cercidiphyllaceae). New Phytol., 196: 617-630. DOI:10.1111/j.1469-8137.2012.04242.x
Qiao, L., Wen, G., Qi, Y., et al., 2018. Evolutionary melting pots and reproductive isolation: a ring-shaped diversification of an odorous frog (Odorrana margaratae) around the Sichuan Basin. Mol. Ecol., 27: 4888-4900. DOI:10.1111/mec.14899
Qiu, Y.X., Fu, C.X., Comes, H.P., 2011. Plant molecular phylogeography in China and adjacent regions: tracing the genetic imprints of Quaternary climate and environmental change in the world's most diverse temperate flora. Mol. Phylogenet. Evol., 59: 225-244. DOI:10.1016/j.ympev.2011.01.012
Qiu, Y.X., Guan, B.C., Fu, C.X., et al., 2009. Did glacials and/or interglacials promote allopatric incipient speciation in East Asian temperate plants? Phylogeographic and coalescent analyses on refugial isolation and divergence in Dysosma versipellis. Mol. Phylogenet. Evol., 51: 281-293. DOI:10.1016/j.ympev.2009.01.016
Qu, J., Liu, N., Bao, X., et al., 2009. Phylogeography of the ring-necked pheasant (Phasianus colchicus) in China. Mol. Phylogenet. Evol., 52: 125-132. DOI:10.1016/j.ympev.2009.03.015
Ray, N., Currat, M., Excoffier, L., 2003. Intra-Deme molecular diversity in spatially expanding populations. Mol. Biol. Evol., 20: 76-86. DOI:10.1093/molbev/msg009
Rogers, A., Harpending, H., 1992. Population growth makes waves in the distribution of pairwise genetic differences. Mol. Biol. Evol., 9: 552-569.
Royden, L.H., Burchfiel, B.C., Van der Hilst, R.D., 2008. The geological evolution of the Tibetan Plateau. Science, 321: 1054-1058. DOI:10.1126/science.1155371
Rozendaal, D.M.A., Hurtado, V.H., Poorter, L., 2006. Plasticity in leaf traits of 38 tropical tree species in response to light; relationships with light demand and adult stature. Funct. Ecol., 20: 207-216. DOI:10.1111/j.1365-2435.2006.01105.x
Schroeder, H., Hoeltken, A.M., Fladung, M., 2012. Differentiation of Populus species using chloroplast single nucleotide polymorphism (SNP) markers - essential for comprehensible and reliable poplar breeding. Plant Biol., 14: 374-381. DOI:10.1111/j.1438-8677.2011.00502.x
Sexton, J.P., Hangartner, S.B., Hoffmann, A.A., 2014. Genetic isolation by environment or distance: which pattern of gene flow is most common?. Evolution, 68: 1-15. DOI:10.1111/evo.12258
Shi, Y.F., Li, J.J., Li, B.Y., et al., 1999. Uplift of the Qinghai-Xizang (Tibetan) Plateau and East Asia environmental change during late Cenozoic. Acta Geograph. Sin., 54: 10-21.
Slatkin, M., 1991. Inbreeding coefficients and coalescence times. Genet. Res., 58: 167-175. DOI:10.1017/S0016672300029827
Soltis, D.E., Gitzendanner, M.A., Strenge, D.D., et al., 1997. Chloroplast DNA intraspecific phylogeography of plants from the Pacific northwest of north America. Plant Syst. Evol., 206: 353-373. DOI:10.1007/BF00987957
Song, X., Milne, R.I., Fan, X., et al., 2021. Blow to the Northeast? Intraspecific differentiation of Populus davidiana suggests a north-eastward skew of a phylogeographic break in East Asia. J. Biogeogr., 48: 187-201. DOI:10.1111/jbi.13992
Spicer, R.A., Farnsworth, A., Su, T., 2020. Cenozoic topography, monsoons and biodiversity conservation within the Tibetan Region: an evolving story. Plant Divers., 42: 229-254. DOI:10.1016/j.pld.2020.06.011
Steig, E.J., 1999. Mid-holocene climate change. Science, 286: 1485-1487. DOI:10.1126/science.286.5444.1485
Sun, Y., Hu, H.Q., Huang, H.W., et al., 2014. Chloroplast diversity and population differentiation of Castanopsis fargesii (Fagaceae): a dominant tree species in evergreen broad-leaved forest of subtropical China. Tree Genet. Genomes, 10: 1531-1539. DOI:10.1007/s11295-014-0776-3
Sun, Y.X., Moore, M.J., Yue, L.L., et al., 2014. Chloroplast phylogeography of the east Asian Arcto-tertiary relict Tetracentron sinense (Trochodendraceae). J. Biogeogr., 41: 1721-1732. DOI:10.1111/jbi.12323
Tamura, K., Peterson, D., Peterson, N., et al., 2011. MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol. Biol. Evol., 28: 2731-2739. DOI:10.1093/molbev/msr121
Tanaka, T., 1954. Species Problem in Citrus: a Critical Study of Wild and Cultivated Units of Citrus, Based upon Field Studies in Their Native Homes. Japanese Society for the Promotion of Science, Tokyo.
Tang, L., Du, C.Q., Jiang, H.L., et al., 2013. Research on gene resources investigation and plus tree selection of Populus lasiocarpa in western Hubei. Hubei For. Sci. Technol., 42: 36-38.
Tian, B., Zhou, Z.L., Du, F.K., et al., 2015. The Tanaka Line shaped the phylogeographic pattern of the cotton tree (Bombax ceiba) in southwest China. Biochem. Syst. Ecol., 60: 150-157. DOI:10.1016/j.bse.2015.04.014
Tian, B., Fu, Y., Milne, R.I., et al., 2020. A complex pattern of post-divergence expansion, contraction, introgression, and asynchronous responses to Pleistocene climate changes in two Dipelta sister species from western China. J. Syst. Evol., 58: 247-262. DOI:10.1111/jse.12524
Usinowicz, J., Chang-Yang, C.H., Chen, Y.Y., et al., 2017. Temporal coexistence mechanisms contribute to the latitudinal gradient in forest diversity. Nature, 550: 105-108. DOI:10.1038/nature24038
Wang, E., Kirby, E., Furlong, K.P., et al., 2012. Two-phase growth of high topography in eastern Tibet during the Cenozoic. Nat. Geosci., 5: 640-645. DOI:10.1038/ngeo1538
Wang, I.J., 2013. Examining the full effects of landscape heterogeneity on spatial genetic variation: a multiple matrix regression approach for quantifying geographic and ecological isolation. Evolution, 67: 3403-3411. DOI:10.1111/evo.12134
Wang, I.J., Bradburd, G.S., 2014. Isolation by environment. Mol. Ecol., 23: 5649-5662. DOI:10.1111/mec.12938
Wang, C.B., Wang, T., Su, Y.J., 2014. Phylogeography of Cephalotaxus oliveri (Cephalotaxaceae) in relation to habitat heterogeneity, physical barriers and the uplift of the Yungui Plateau. Mol. Phylogenet. Evol., 80: 205-216. DOI:10.1016/j.ympev.2014.08.015
Wang, P., Scherler, D., Jing, L.Z., et al., 2014. Tectonic control of Yarlung Tsangpo Gorge revealed by a buried canyon in southern Tibet. Science, 346: 978-981. DOI:10.1126/science.1259041
Wang, P., Yao, H., Gilbert, K.J., et al., 2018. Glaciation-based isolation contributed to speciation in a Palearctic alpine biodiversity hotspot: evidence from endemic species. Mol. Phylogenet. Evol., 129: 315-324. DOI:10.1016/j.ympev.2018.09.006
Wang, W.M., 1994. Paleofloristic and paleoclimatic implications of Neogene Palynofloras in China. Rev. Palaeobot. Palynol., 82: 239-250. DOI:10.1016/0034-6667(94)90078-7
Wang, Y.H., Comes, H.P., Cao, Y.N., et al., 2017. Quaternary climate change drives allo-peripatric speciation and refugial divergence in the Dysosma versipellis-pleiantha complex from different forest types in China. Sci. Rep., 7: 40261. DOI:10.1038/srep40261
Wang, M., Zhang, L., Zhang, Z., et al., 2020. Phylogenomics of the genus Populus reveals extensive interspecific gene flow and balancing selection. New Phytol., 225: 1370-1382. DOI:10.1111/nph.16215
Wang, Y.Q., Feijó, A., Cheng, J.L., et al., 2021. Ring distribution patterns—diversification or speciation? Comparative phylogeography of two small mammals in the mountains surrounding the Sichuan Basin. Mol. Ecol., 30: 2641-2658. DOI:10.1111/mec.15913
Wen, Z.X., Wu, Y., Ge, D.Y., et al., 2017. Heterogeneous distributional responses to climate warming: evidence from rodents along a subtropical elevational gradient. BMC Ecology, 17: 17. DOI:10.1186/s12898-017-0128-x
Wu, Z.Y., Wu, S.G., 1996. A proposal for a new floristic kingdom (realm) -the E. Asiatic kingdom, its delimitation and characteristics - ScienceOpen. In: Presented at the Proceedings of the First International Symposium on Floristic Characteristics and Diversity of East Asian Plants. Higher Education Press, Beijing.
Xiao, J.H., Ding, X., Li, L., et al., 2020. Miocene diversification of a golden-thread nanmu tree species (Phoebe zhennan, Lauraceae) around the Sichuan Basin shaped by the East Asian monsoon. Ecol. Evol., 10: 10543-10557. DOI:10.1002/ece3.6710
Xie, X.F., Yan, H.F., Wang, F.Y., et al., 2012. Chloroplast DNA phylogeography of Primula ovalifolia in central and adjacent southwestern China: past gradual expansion and geographical isolation. J. Syst. Evol., 50: 284-294. DOI:10.1111/j.1759-6831.2012.00204.x
Xu, J., Deng, M., Jiang, X.L., et al., 2015. Phylogeography of Quercus glauca (Fagaceae), a dominant tree of East Asian subtropical evergreen forests, based on three chloroplast DNA interspace sequences. Tree Genet. Genomes, 11: 805. DOI:10.1007/s00269-015-0764-7
Yan, H.F., Zhang, C.Y., Wang, F.Y., et al., 2012. Population expanding with the Phalanx model and lineages split by environmental heterogeneity: a case study of Primula obconica in Subtropical China. PLoS One, 7: e41315. DOI:10.1371/journal.pone.0041315
Yang, A., Zhong, Y., Liu, S., et al., 2019. New insight into the phylogeographic pattern of Liriodendron chinense (Magnoliaceae) revealed by chloroplast DNA: east–west lineage split and genetic mixture within western subtropical China. PeerJ, 7: e6355. DOI:10.7717/peerj.6355
Ye, J.W., Zhang, Y., Wang, X.J., et al., 2017. Phylogeographic breaks and the mechanisms of their formation in the Sino-Japanese floristic region. Chin. J. Plant Ecol., 41: 1003-1019. DOI:10.17521/cjpe.2016.0388
Ying, L.X., Zhang, T.T., Chiu, C.A., et al., 2016. The phylogeography of Fagus hayatae (Fagaceae): genetic isolation among populations. Ecol. Evol., 6: 2805-2816. DOI:10.1002/ece3.2042
Young, N.D., Healy, J., 2003. GapCoder automates the use of indel characters in phylogenetic analysis. BMC Bioinformatics, 4: 6. DOI:10.1186/1471-2105-4-6
Zhang, L., Xi, Z.X., Wang, M.C., et al., 2018. Plastome phylogeny and lineage diversification of Salicaceae with focus on poplars and willows. Ecol. Evol., 8: 7817-7823. DOI:10.1002/ece3.4261
Zhao, Y.J., Gong, X., 2015. Genetic divergence and phylogeographic history of two closely related species (Leucomeris decora and Nouelia insignis) across the 'Tanaka Line' in Southwest China. BMC Evol. Biol., 15: 134. DOI:10.1186/s12862-015-0374-5
Zhao, Y.M., Zhang, L., 2015. The phylogeographic history of the self-pollinated herb Tacca chantrieri (Dioscoreaceae) in the tropics of mainland Southeast Asia. Biochem. Syst. Ecol., 58: 139-148. DOI:10.1016/j.bse.2014.11.011
Zhou, S.Z., Wang, X.L., Wang, J., et al., 2006. A preliminary study on timing of the oldest Pleistocene glaciation in Qinghai-Tibetan Plateau. Quat. Int., 154–155: 44-51.
Zhou, W., Yan, F., Fu, J., et al., 2013. River islands, refugia and genetic structuring in the endemic brown frog Rana kukunoris (Anura, Ranidae) of the Qinghai-Tibetan Plateau. Mol. Ecol., 22: 130-142. DOI:10.1111/mec.12087