Skip to main content

Population differentiation and dynamics of five pioneer species of Gaultheria from the secondary forests in subtropical China

Abstract

Background

The influence of native secondary succession associated with anthropogenic disturbance on the biodiversity of the forests in subtropical China remains uncertain. In particular, the evolutionary response of small understory shrubs, particularly pioneer species inhabiting continuously disturbed habitats, to topographic heterogeneity and climate change is poorly understood. This study aimed to address this knowledge gap by focusing on the Gaultheria crenulata group, a clade of small pioneer shrubs in subtropical China.

Results

We examined the genetic structure and demographic history of all five species of the G. crenulata group with two maternally inherited chloroplast DNA (cpDNA) fragments and two biparentally inherited low-copy nuclear genes (LCG) over 89 natural populations. We found that the genetic differentiation of this group was influenced by the geomorphological boundary between different regions of China in association with Quaternary climatic events. Despite low overall genetic diversity, we observed an isolation-by-distance (IBD) pattern at a regional scale, rather than isolation-by-environment (IBE), which was attributed to ongoing human disturbance in the region.

Conclusion

Our findings suggest that the genetic structure of the G. crenulata group reflects the interplay of geological topography, historical climates, and anthropogenic disturbance during the Pliocene–Pleistocene-Holocene periods in subtropical China. The observed IBD pattern, particularly prominent in western China, highlights the role of limited dispersal and gene flow, possibly influenced by physical barriers or decreased connectivity over geographic distance. Furthermore, the east-to-west trend of gene flow, potentially facilitated by the East Asian monsoon system, underscores the complex interplay of biotic and abiotic factors shaping the genetic dynamics of pioneer species in subtropical China’s secondary forests. These findings can be used to assess the impact of environmental changes on the adaptation and persistence of biodiversity in subtropical forest ecosystems.

Peer Review reports

Background

Subtropical China, is recognized as one of the world’s most species-rich forested regions [1]. Its ecosystems significantly contribute to biodiversity preservation and helps to maintain equilibrium of the global carbon cycle [2]. Occupying the region south of the Qinling-Huaihe River Line, north of the Leizhou Peninsula, and east of the Hengduan Mountains (i.e., 22–34°N, 98°E to the coastline in the east [3]), this area encompasses the main southern parts of China’s second-step and third-step terrain regions. The second step is characterized by plateaus and basins ranging from 1,000 to 2,000 m above sea level (a.s.l.), whereas the third-step comprises plains and hills below 500 m a.s.l. [4]. The forests in subtropical China have a highly heterogeneous topography and complex environmental history, with intense mountain uplifts and Pliocene–Pleistocene climatic dynamics, harboring refugia for many ancient lineages [5].

Subtropical China’s forests have undergone significant transformation into plantations and secondary forests due to prolonged and persistent anthropogenic disturbance, with some areas being replaced by densely populated agricultural lands [6, 7]. The influence of these secondary forests on the biodiversity of subtropical China has been a topic of long-standing debate in community ecology [8, 9]. They have been posited to decrease biodiversity [10], yet also have been thought to serve as biodiversity reservoirs [11]. Research on these forests has predominantly focused on ecosystem stability, species composition, and community/stand structure for ecosystem restoration [10, 12]. Less emphasis has been placed on adaptation to disturbance and successional dynamics of pioneer species in this ecosystem, particularly within the context of phylogeography at a geological scale. Secondary succession can affect genetic diversity [13], and so studying the population genetic dynamics of pioneer plant species may provide valuable insights into succession [14].

Population genetic studies on pioneer species in the secondary forests of the subtropical China have concentrated on widespread and dominant trees or large shrubs such as Quercus fabri Hance [15] and Liquidambar formosana Hance [16]. In contrast, short/small pioneer shrub species of the forest understory in subtropical China have largely been neglected. Although small understory shrubs are not the dominant element of the forests in subtropical China, their functional traits, activities, succession rate, and phylogenetic diversity in areas under secondary succession have been found to be influenced more strongly by topographic heterogeneity or microhabitats (i.e., more likely to escape extreme cold) than by changing climates [5, 17,18,19]. Conversely, other studies have found that such shrubs are strongly responsive to recent climate change [20, 21].

The Gaultheria crenulata group is a clade of five species of small understory shrubs, including G. crenulata Kurz, G. luchunensis Yi R. Li, Lu Lu & P.W. Fritsch, G. pingbienensis (C.Y. Wu ex T.Z. Xu) Yi R. Li, Lu Lu & P.W. Fritsch, G. wuliangshanensis Yi R. Li, Lu Lu & P.W. Fritsch, and G. mangshanensis Yi R. Li, Lu Lu & P.W. Fritsch [22]. The species of the group are some of the earliest pioneer colonizers in the secondary coniferous mixed forests in subtropical China. G. crenulata is the most widespread species of Chinese Gaultheria, occurring nearly throughout the area of China south of the Yangtze River Basin, and ranging from 200 to 3300 m elevation [22]. Because of the rich presence of methyl salicylate (oil of wintergreen; [23], this species is used in Chinese traditional medicine by > ten minorities [24] and which now is suffering heavy local harvesting. This is leading to the reduction of populations or individuals, as observed by us in the field. The G. crenulata group diverged earlier in the phylogenetic history of Gaultheria than most other Chinese species of the genus, and its crown age is estimated as ca. 4.1 Ma in the early Pliocene [25]. Thus, the genetic differentiation and demographic dynamics of the group can be expected to provide insights into the impacts of topography and climates from the Pliocene to present time on species adaptations in subtropical China. This in turn will aid understanding as to how these pioneer plants evolved in current secondary forests.

Here we combined geographic, environmental, and genetic data with a phylogeographic approach to uncover the genetic structure and dynamics of the Gaultheria crenulata group. We address the following questions on the secondary forests of the subtropical China. (1) In subtropical China, what is the pattern and strength of genetic differentiation of the group? (2) How did the group respond to long-term climatic and topographic dynamics? Does isolation by distance (IBD) or isolation by environment (IBE) better explain its genetic differentiation? (3) How have variations in topography and climate through time influenced the genetic differentiation and demographic history of the group?

Results

Genetic diversity and differentiation statistics

For the data from low-copy nuclear genes (LCG), the combined sequence totaled 1,442 bp in length, with AAT(aspartate aminotransferase) 670 bp and LOC (far1-related region) 772 bp. We identified 292 polymorphic sites, including 10 singleton variable sites and 282 parsimony-informative sites. The allelic richness (AR) of the whole dataset was 542 and nucleotide diversity (π) was 7.67 × 10–3. Genetic variation differs among populations, with π ranging from 0.00 to 9.95 × 10–3 and AR ranging from 1 to 1.98 (Table 1). The highest level of genetic diversity was found in the SB population (AR = 1.97, π = 9.95 × 10–3). Significant genealogically geographic structure was found in the overall (population) dataset (NST = 0.35 > GST = 0.19; P < 0.01). The combined length of the two chloroplast DNA (cpDNA) fragments was 1,268 bp, comprising rpl33-psaJ at 669 bp and trnL-rpl32 at 599 bp. We identified 29 polymorphic sites, consisting of 11 singleton variable sites and 18 parsimony-informative sites. The concatenation analysis based on cpDNA data revealed an overall haplotype diversity (Hd) of 0.31 and nucleotide diversity ) of 1.26 × 10–3 across 89 populations. Based on the PERMUT analysis of cpDNA data, the Gaultheria crenulata group revealed low to moderate genetic variation among populations at the species level (HT = 0.34) and low genetic variation within populations (HS = 0.091).Genetic differentiation incorporating distances among haplotypes (NST = 0.813) was found to be significantly higher than population differentiation without incorporating distance (GST = 0.739, P < 0.05).

Table 1 Sample location, sample size and genetic diversity of two cpDNA and two LCG (low-copy nuclear genes) loci, and cpDNA haplotypes in 89 sampled populations of the G. crenulata group. Hd, cpDNA haplotype richness; π, nucleotide diversity; AR, nuclear sequence allelic richness

Population genetic structure and migration

For the LCG data, Bayesian clustering analysis with STRUCTURE indicated that the sequential increase in K values from 2 to 10 (see Fig. S1) provided into the relative strength of different signals of population subdivision within the G. crenulata group. The ΔK value was highest when K = 2, followed by K = 8, K = 3, and K = 5. In our analysis at K = 2, the data from the 89 populations revealed a significant divide along the Wushan and Xuefeng Mountains, which form the boundary between the second and third-step regions of China (Fig. 1). The western China group (WC, 60 populations) and eastern China group (EC, 29 populations) were genetically divided and exhibited a mosaic of clusters in the WC group (Table 1; Fig. 1). When K = 3, the results were consistent with K = 2. At K = 5, G. wuliangshanensis (JD in Fig. S1) and G. luchunensis (YLC in Fig. S1) separated from the WC group, forming a separate subpopulation, whereas G. mangshanensis (NMS in Fig. S1) separated from the EC group, forming another distinct subpopulation. At K = 8, G. pingbienensis separated from the WC group, forming a new subpopulation (DWS in Fig. S1). PCoA revealed distinct genetic structure among populations within the G. crenulata group, with the first and second axes capturing 27.6% and 16.1% of the total genetic variation, respectively. The WC group (shown in red) and EC group (shown in blue) exhibited distinct clusters, indicating their differentiation (Fig. 1).

Fig. 1
figure 1

a Geographical distribution of the genetic clusters in 89 populations sampled from the G. crenulata group in subtropical China. Two genetic clusters are generated from LCG data (red/blue, and blue) with STRUCTURE analysis, K = 2 (see population codes in Table 1). The black dashed line represents the boundary that generally divides into two clusters (the WC group and the EC group), which also roughly corresponds to the boundary between the second-step region and the third-step region of China’s terrain; b Principal component analysis of LCG data of 1,435 individuals from 89 populations of the two lineages; c Admixture assignment for 89 populations with K = 2: each bar represents a population, with different colors corresponding to one of the ancestries, which was determined to be the optimal K value with the ΔK method. (Map source: https://www.webmap.cn/main.do?method=index)

Between the two cpDNA fragments combined, 23 haplotypes were obtained, with haplotypes H1 (in central and northern Yunnan and southern Sichuan) and H2 (in the remaining regions) accounting for two major components (Fig. 2). In the network analysis, 23 chloroplast haplotypes separated into two major groups. The first group consisted of 12 haplotypes (H3, H6, H7, H10, H11, H16, H17, H18, H21, H22, and H23) with H1 as the center, which then diverged in a star-like pattern to form a lineage. The second group comprised 11 haplotypes (H4, H5, H8, H9, H12, H13, H14, H15, H19, and H20) with H2 as the center, which also differentiated in a star-like pattern to form another lineage (Fig. 2). The SplitsTree analysis recovered G. wuliangshanensis, G. pingbienensis, G. luchunensis, and G. mangshanensis as forming sister groups distinct from G. crenulata (Fig. S2). The Bayesian search strategy-based analyses with MIGRATE for the LCG data indicated that long-term gene flow between populations is asymmetric, with a greater east-to-west tendency (EC group Θi = 0.097 vs. WC group Θi = 0.006; Nm (EC to WC) = 88.87 >  > Nm (WC to EC) = 2.94; S. M. Table S1).

Fig. 2
figure 2

a Geographic distribution of the cpDNA haplotypes in 89 populations sampled from the G. crenulata group in subtropical China (see population codes in Table 1); b Network plot of the 23 cpDNA haplotypes recovered, where each haplotype is depicted with the same color as in the geographic distribution. The size of each circle corresponds to the frequency of each haplotype. The red diamonds on the branches represent hypothetical missing haplotypes. (Map source: https://www.webmap.cn/main.do?method=index)

In the AMOVA analyses, for the LCG data 14.45% of the variation was found between the WC group and the EC group. For each of the two groups, 27.0% was attributed to among-populations variation and 58.6% within-population variation (FCT = 0.144, FST = 0.414, P < 0.01; Table 2). For cpDNA data, 5.5% of the variation was found between the two groups, 75.8% among populations within a group, and 18.7% within populations (FCT = 0.055, FST = 0.814, P < 0.01; Table 2).

Table 2 Analysis of molecular variance (AMOVA) within the G. crenulata group based on cpDNA and LCG data

Demographic history and divergence time estimates

For the LCG data, the expansion model was not rejected by either the SSD or the HRag statistics (P > 0.05, Table S2). The neutral tests revealed an overall selection effect across populations (Tajima’s D =  − 1.8803, P < 0.05; Fu and Li’s D* = 3.7249, P < 0.05). With the two groups analyzed separately, only the WC group exhibited significant values (Tajima’s D =  − 1.7758, P < 0.05; Fu and Li’s D* = 3.4592, P < 0.05), and mismatch analysis supported the historical population expansion model (both P > 0.05; Table S2). For cpDNA data, both the SSD and HRag indices did not reject the expansion model (P > 0.05; Table S2). However, the neutral test for the total populations showed significance (Tajima’s D =  − 1.3975, P < 0.05; Fu and Li’s D* =  − 3.8003, P < 0.05). With the two groups analyzed separately, the WC group exhibited a non-significant value of Tajima’s D =  − 0.9904 (P > 0.05). However, the value of Fu and Li’s D* =  − 4.1559 (P < 0.05) was significant, suggesting population expansion. The neutral tests of the EC group were insignificant (Tajima’s D =  − 1.2438, P > 0.05; Fu and Li’s D* = 1.2038, P > 0.05; Table S2).

The DIYABC Random Forest analysis provided clear evidence of admixture within the Gaultheria crenulata group (Fig. 3). When the ten scenarios are analyzed separately, the observed data fall approximately within the range of the output of the conformity model on Axis 1 but fall outside of the range on Axis 2 (Fig. 3). The highest classification votes and estimated posterior probabilities showed that scenario 9, i.e., the ancient differentiation with recent mixing model, has the highest posterior probability (0.444) among all hypothetical scenarios (Table S3-S5). This suggests that the WC and EC groups diverged from a most recent common ancestor (NA). The G. mangshanensis population was identified as a genetically admixed population of the WC and EC groups. The time of divergence between the WC group and the EC group was estimated as 0.617 Ma. Demographic bottlenecks occurred ca. 5,340 generations ago (corresponding to roughly 26.7 kya, with the average five-year reproductive maturity period for Gaultheria considered [26]. These bottleneck events occurred prior to the divergence of G. pingbienensis (DWS), G. luchunensis (YLC), G. wuliangshanensis (JD), and G. mangshanensis (NMS) during the interval from 0.103 Ma to 0.140 Ma (Tables S3-S6, Fig. 3).

Fig. 3
figure 3

a Projection of the LCG dataset on a single LDA axis for the training set when analyzing the two hypotheses of scenarios (nonadmixed and admixture); b Projection of the LCG dataset on the first two LDA axes when analyzing the ten scenarios separately. Each colored point corresponds to a single simulation of one of the models; the red dot corresponds to the observed data. The observed data fall approximately within the range of the output of the conformity model on Axis 1 but fall outside of the range on Axis 2; c The best demographic model of divergence for the six classified groups in the G. crenulata group. Each block represents a current or ancestral population and its estimated effective population size. Arrows indicate the direction of gene flow, and the estimated migration rate is marked above or below the arrow. The timing of the splitting events is indicated in million years ago (Ma). The admixture ratio is indicated by the additional parameters “ra” and “1-ra,” representing the genetic contribution of the ancestral population Pop 1 and Pop 2, respectively

From the BEAST results of cpDNA haplotypes, the estimated crown age of the G. crenulata group was inferred as early Pliocene (4.06 Ma, 95% HPD: 1.98–5.93 Ma, PP = 1.00). The haplotypes form two major subclades (C1 and C2; Fig. 4). The divergence time estimate for subclade C1 is 3.16 Ma (95% HPD: 1.31–5.05 Ma, PP = 0.99). This subclade is composed of haplotypes from most populations of the WC group and two of the EC group (NMS of G. mangshanensis and DWS, haplotypes H10 and H11, with divergence at ca. 1.2 Ma, 95% HPD: 0.12–2.59 Ma, PP = 1.00). The divergence time estimate for subclade C2 was 3.28 Ma (95% HPD: 1.48–5.21 Ma, PP = 0.98). This subclade is composed of haplotypes from most populations of the EC group and all private haplotypes. The private haplotypes H22 and H23 of G. luchunensis (YLC) form a sister group that diverged ca. 1.45 Ma (95% HPD: 0.11–3.03 Ma, PP = 1.00). Haplotype H4, exclusive to G. pingbienensis (DWS), diverged ca. 1.17 Ma (95% HPD: 0.04–2.65 Ma, PP = 0.85).

Fig. 4
figure 4

Chronogram of the G. crenulata group based on cpDNA data. The times in both best ages (in bold) and their 95% HPD with timescale blue bars (font in smaller size within brackets) are presented above the branches and posterior probability below the branches (only the divergence times for the nodes with posterior probability > 50% are shown)

IBD and IBE analyses

The MMRR analysis and Mantel analysis for both the LCG and cpDNA data showed a significant IBD (P < 0.01) and non-significant IBE (P > 0.05) across the overall (populations). When the WC group and EC group were tested separately, a positive IBD correlation was detected only in the WC group (P < 0.01; Table 3).

Table 3 Multiple matrix regression with randomization (MMRR) analysis and Mantel test results showing the relative contribution of isolation by spatial distance (IBD) and environment (IBE) with respect to multiple population genetic distance indices of six datasets: 1. LCG data of the overall populations, 2. cpDNA data of the overall populations, 3. LCG data of populations from the western China group, 4. cpDNA data of populations from the western China group, 5. LCG data of populations from the eastern China group, 6. cpDNA data of populations from the eastern China group

Ecological niche modeling

The average test AUC for replicate runs was 0.968 (SD = 0.010) for the WC group and 0.988 (SD = 0.004) for the EC group, indicating satisfactory model performance. The jackknife test revealed that the variable Bio14 (Precipitation of Driest Month) contributed most to the potential distribution modeling of the WC group (33.9%) and EC group both (68.3%), followed by Bio1 (Annual Mean Temperature) at 20.0% of the WC group and Bio5 (Max Temperature of Warmest Month) at 8.9% of the EC group (Table S7). The current distribution projections accurately depict the optimal areas for the G. crenulata group in the subtropical mountainous regions of China, spanning from northwestern to southeastern Yunnan, Guizhou, the border regions of Hunan, Guangdong, and Guangxi, and southeastern Fujian. During the LIG and MH periods, the potential range was slightly larger but less optimal in northwestern Yunnan. The WC group’s potential range during these periods was larger but also less optimal in northwestern Yunnan and Southern Sichuan. For the EC group, the potential range during the MH period was relatively smaller and more scattered, expanding from then to the present. Under the maximum greenhouse gas emission model of RCP 8.5 scenario, the overall geographic distribution is expected to shrink, with a significant reduction in optimal areas such as Chongqing, Hubei, and Northern Guizhou of the WC group (Fig. 5). The EC group, which is more heavily affected by anthropogenic disturbances, exhibits a more limited and fragmented optimal distribution area (Fig. 6).

Fig. 5
figure 5

The predicted distribution of the WC group of the G. crenulata group in each of four climate scenarios A–D. a Last glacial maximum (LGM) based on the output of CCSM4; b Mid-Holocene (MH) based on the output of CCSM4; c present (1950–2000) based on the output of CCSM4; d future (2070) based on the output of CCSM4. (Map source: https://www.webmap.cn/main.do?method=index)

Fig. 6
figure 6

The predicted distribution of the EC group of the G. crenulata group in each of four climate scenarios A–D. a Last glacial maximum (LGM) based on the output of CCSM4; b Mid-Holocene (MH) based on the output of CCSM4; c present (1950–2000) based on the output of CCSM4; d future (2070) based on the output of CCSM4. (Map source: https://www.webmap.cn/main.do?method=index)

Discussion

The biogeographic boundaries and chloroplast-nuclear discordance

Our LCG data revealed a significant divide along the Wushan and Xuefeng Mountains, which coincides with the boundary between the second- and third-step regions of China. This finding aligns with the east–west differentiation observed in many plant groups in subtropical China, e.g., Quercus acutissima Carruth. and Toxicodendron vernicifluum (Stokes) F. Barkley [4, 27]. However, for cpDNA data, a biogeographic division of the G. crenulata group into two lineages was observed: one from the high-elevation region within the second-step region of China (reaching nearly 2,000 m) in central and northern Yunnan and southern Sichuan, mainly characterized by subalpine mountains and mid-elevation basins with fragmented regions isolated by valleys (e.g., Jinsha River, Dadu River, Yalong River, and Anning River), the other from the lower-elevation area (ca. 1,000 m), characterized by gentle undulating hills and covering most of the eastern Yungui Plateau and the middle-lower reaches of the Yangtze River in the third-step plain (Fig. 2).

This contrast between nuclear and cpDNA data serves as evidence of how different genomes respond to past population fluctuations and biogeographic events [28] and highlights the influence of historical factors and geographical barriers in shaping the genetic structure of the G. crenulata group. The biogeographic nuclear-chloroplast discordance detected here has been observed in various other plant species. For example, mountain barriers led to chloroplast differentiation whereas nuclear gene flow remained extensive in Ulmus lamellosa C.Wang & S.L.Chang [29]. The distinct geographic structure between nDNA and cpDNA data for Quercus fabri is evident, facilitating effective nuclear gene flow through pollen [30]. The chloroplast-nuclear discordance exhibited in our study may be attributed to complex topography and limited distance of cytoplasmic gene flow via seed dispersal, and widespread recombination of the nuclear genome across the Southwest China indicating effective nuclear gene flow via pollen [31].

Possible reasons for the genealogical differentiation of the chloroplast data of populations near the west boundary of Yunnan possessing haplotypes of the east group of the G. crenulata group are firstly, that it corresponds with the “Tanaka line” identified by Tanaka in 1954 [32], indicating a split in the distribution of Citrus reticulata Blanco in Yunnan. This division likely influenced the early differentiation of the G. crenulata group. Secondly, changes in major water systems through the uplift of the Tibetan Plateau, such as diversions may also be relevant [33, 34]. For instance, the upper Jinsha, Yalong, and Dadu rivers, which once flowed into the ancient Red River, now exhibit a discontinuous structure. Geological events, roughly between the Miocene and Pleistocene, caused these alterations [35, 36], coinciding with the time of division of the G. crenulata group into two lineages (4.06 Ma, 95% HPD: 1.98–5.93 Ma, PP = 1.00). This suggests that the group was once continuously distributed in the valleys of ancient rivers, but present-day patterns hinder chloroplast gene flow between populations, which has led to significant genetic differentiation. Moreover, rivers like the Lancang, Nu, and Hong have maintained high levels of population dispersal and gene flow [36]. Similar patterns influenced by riverine changes are observed in other species such as Terminalia franchetii Gagnep., Schizopygopsis malacanthus Herzenstein, and Incarvillea arguta [37,38,39].

The role of IBD in shaping the genetic structure of the G. crenulata group

We found that IBD rather than IBE influenced the genetic differentiation of the G. crenulata group, with a positive IBD correlation in the WC group as supported by both LCG and cpDNA data. Gene flow may follow patterns of IBE, whereby rates of gene flow rates higher among similar environments [40]. An IBD pattern is also often observed when gene flow is inversely proportional to geographic distance, limiting genetic variation in response to selection pressures primarily through natural dispersal and population size [41]. Genetic differentiation by IBD rather than IBE has been detected in many widespread species (e.g., Echinacea angustifolia DC [42], and Mimulus guttatus DC [43]), and this pattern has been suggested to facilitate colonization of new areas under conditions of constant disturbance [41, 44]. Because G. crenulata is a pioneer species, it can establish itself in newly created or recently disturbed environments after deforestation. These environments are more prone to continuous expansion due to human-induced forest degradation in subtropical China. This ongoing expansion process is likely to be driven more by geographic distance, resulting in genetic differentiation (IBD) rather than environmental adaptation (IBE).

We inferred a variant of the IBD pattern at a regional scale: the IBD pattern was only observed in the WC group and not in the EC group. IBD is often detected in populations of plants that tend to move shorter distances, resulting in a pattern of genetic divergence that increases significantly with geographic distance [45, 46]. In contrast to Eastern China, the topography of Western China is characterized by higher heterogeneity (e.g., a wider elevational range), facilitating genetic differentiation associated with phenological differences [47]. It is also characterized by more complex topographic barriers (e.g., high mountains and deep river valleys [36]), which, in the case of the G. crenulata group, have restricted dispersal and gene flow of the WC group. Accordingly, the absence of an IBD and IBE pattern in the EC group suggests higher gene flow and habitat homogeneity, possibly from recent colonization or extinction events and/or historical or biogeographic migration [41].

Demographic history of the G. crenulata group

Our data revealed a population expansion in the WC group but not in the EC group. Ecological niche modeling indicated that the optimal potential range of the WC group (i.e., northwestern Yunnan and Southern Sichuan) had expanded from the MH period (ca. 6 kya) within the warm-period Holocene Thermal Maximum (5–11 kya). This period is characterized as a transition from arid to wet conditions [48, 49]. A similar expansion event occurring in Western China has been inferred in Fokienia hodginsii (Dunn) A.Henry & H.H.Thomas [50] and Torreya fargesii Franch. [51], the effects of glacial cycles could have been strongest in this area. The Hengduan Mountains serve as a migration corridor and refuge, facilitating rapid forest migration and providing safe havens during extreme climate events. As conditions improve, species can expand beyond these refuges [52]. Both climate-driven Holocene and geographical refugia facilitated the expansion of the WC group in this study. The ecological niche model also predicts contraction of the current optimal range of the G. crenulata group and a northward shift of its suboptimal range, strongly influenced by precipitation of the driest month, temperature annual range, and mean temperature of the wettest quarter. This potential range shift is consistent with the general trend of species shifting towards higher latitudes through global warming [53].

Based on the clustering results by STRUCTURE, the spatial genetic structure displayed a mosaic of clusters in the WC group. The mosaic population structure has been thought to likely result from stochastic short-range dispersal of individuals [54], also indicated by IBD detected in this group. Our study revealed historical gene flow of the G. crenulata group from east to west based on the LCG data (Nm (EC to WC) = 88.87 >  > Nm (WC to EC) = 2.94). Asymmetrical gene flow from east to west was also found as occurring in widespread Chinese species such as Juglans cathayensis Dode [55], Arnebia szechenyi Kanitz [56], and Q. acutissima  [57]. This asymmetry was postulated as facilitated by the East Asian monsoon, hybrid incompatibility, and wind-mediated pollen dispersal. Species of Gaultheria can be pollinated by insects and wind [58, 59]. In our study, east-to-west gene flow of the G. crenulata group may be enhanced by pollen-mediated movement influenced by the East Asian monsoon system, the significant role of which in shaping population genetic structure has also been demonstrated in plants such as Primulina Hance [60] and Quercus chenii Nakai [61].

According to the DIYABC-RF analysis, the narrowly endemic species G. pingbienensis (DWS), G. luchunensis (YLC), G. wuliangshanensis (JD), and G. mangshanensis (NMS) were inferred to have diverged from their ancestral population ca. 0.1 (0.103–0.140) million years ago during the Pleistocene refugium period (Fig. 3-C). Previous studies have reported the existence of multiple refugia during the Quaternary ice age in this region [62]. The presence of these refugia facilitated the divergence of surviving species through long-distance isolation [12, 63]. Additionally, we found that distinct haplotypes of these four species successively emerged during the Pleistocene (Fig. 4), indicating significant divergence events during Pleistocene glaciation. These four species thrive in undisturbed high mountain ridges and moist slope forest environments with high endemism. Gaultheria pingbienensis, G. luchunensis, G. wuliangshanensis are distributed in the Wuliang Mountains of the Central Yunnan Plateau, and G. mangshanensis is distributed in the Nanling Mountains of the Chinese southern hills. Conversely, G. crenulata prefers disturbed secondary forests with full or partial sun exposure [22]. The isolation of populations on separate mountains, which occurred during warm periods when the species followed its optimal climate towards higher elevations, was probably caused by ancient glacial cycles, which occurred from over 1 Ma to about 0.1 Ma BP [64]. Related studies found that during the last interglacial period (0.1 Ma), summer monsoon rains and temperatures increased in vast areas across Asia [65], which may have also facilitated the divergence of these four species. Furthermore, the isolation of G. crenulata of the WC group and EC group at 0.617 Ma BP, although more recent compared to the uplift of the Himalayan–Tibetan Plateau and the formation of the staircase topography since the Late Miocene times [66], has been detected in many other species as well. For example, the divergence time between the eastern and western lineages of Liriodendron chinense (Hemsl.) Sarg is estimated to be ca. 0.433 Ma [67] and Asteropyrum J. R. Drumm. & Hutch. is dated to 1.2 Ma [68]. Our data indicate that the east–west lineage split is a gradual process of geographic isolation and allopatric speciation.

Population genetic differentiation and conservation in the secondary forests of the subtropical China

The significant genetic differentiation observed in the Gaultheria crenulata group along an east–west axis in subtropical China is consistent with patterns seen in other tree species in the same region, such as Quercus acutissima  [4]., Q. fabri [15], and L. chinense  [69]. This differentiation is primarily attributed to historical geographic and climatic shifts during the Pliocene. Conversely, tree species in the secondary forests, such as Camellia oleifera Abel and Liquidambar formosana Hance, did not display genetic differentiation along an east–west axis [70]. Instead, they showed low genetic diversity within populations and high genetic differentiation among populations, largely due to anthropogenic disturbance and habitat fragmentation [16, 70]. Anthropogenic activities, such as deforestation for agriculture, urbanization, and logging, have led to habitat loss and fragmentation [71], gexacerbating genetic isolation among populations [72]. The conversion of natural forests into plantations and secondary forests has created barriers to gene flow, resulting in genetic differentiation among isolated populations of the pioneer tree Cecropia hololeuca Miq. [73]. Human activities, such as changes in population density, cropland use, and irrigated rice area, are significantly linked to the demographic history of Litsea elongata (Wall. ex Ness) Benth. & Hook. f [74].These disturbances, combined with climate change, pose significant threats, including reduced habitat range, increased risk of extinction, and loss of genetic diversity [75].

In our study, the shrubs of the G. crenulata group exhibited low genetic variation within populations (HS = 0.091) and moderate levels of genetic variation among populations (HT = 0.34). This indicates that although there is some genetic diversity within populations, there is also significant genetic differentiation among populations. Similarly, Pinus armandii Franch, a companion tree species to the G. crenulata group, showed lower levels of genetic diversity in planted populations, highlighting the broader challenge of maintaining genetic diversity in the plant species of the region [76].

Our field observations indicate declines in populations, particularly in regions with frequent specimen records from the past, such as Mount Emei in Sichuan Province. This decline may be attributed to fluctuations in temperature and precipitation over the past decades [77]. Additionally, populations of the G. crenulata group are usually clonal, reproducing vegetatively through rhizomes [78]. The populations in montane central and northeastern Yunnan have been removed by local minorities or pharmacy companies which has also substantially reduced the biomass of this medicinal plant.

The significant genetic differentiation observed in the G. crenulata group, influenced by both historical factors and contemporary anthropogenic disturbance, underscores the importance of conservation efforts to preserve this species. Anthropogenic activities, including deforestation, urbanization, and habitat fragmentation, have contributed to the genetic differentiation and decline of these species. Our study predicts that as climate warms, the potential distribution of the G. crenulata group will continuously shrink, especially in optimal areas like Chongqing, Hubei, and Northern Guizhou (Fig. 5). Thus, conservation strategies should address the complex interplay of historical and environmental factors, as well as anthropogenic activities, to ensure the genetic diversity and future survival of the G. crenulata group.

Conclusions

The low genetic diversity and IBD pattern detected in the WC group of G. crenulata for gene differentiation can be inferred to result from the disturbance of human activities. The region covering central and northern Yunnan and southern Sichuan is characterized by the potentially optimal range with highest genetic diversity of the G. crenulata group. Such a region possesses high ability to withstand adverse environmental conditions and plays a vital role in promoting ecosystem restoration [79]. From a protection perspective, it is crucial to enhance conservative efforts in this region and promote habitat restoration, to ensure sustainable utilization of the G. crenulata group.

Materials and methods

Field sampling

The sampling design encompassed the entire geographic range of the Gaultheria crenulata group in subtropical China based on a recent taxonomic revision for the group [22]. To fully cover the geographic extent of our focal group, we consulted nearly all digitized specimens of the group in the Chinese Virtual Herbarium Database (https://www.cvh.ac.cn/), with a total of 1,434 specimens (1924–2016). Based on the locality information from these specimens, we collected 1,435 individuals from 89 natural populations (Fig. 1), including 85 from G. crenulata (widely distributed across the entirety of the subtropical China), one (YLC) from G. luchunensis (Luchun, Yunnan), one (DWS) from G. pingbienensis (Pingbian, Yunnan), one (JD) from G. wuliangshanensis (Jingdong, Yunnan), and one (NMS) from G. mangshanensis (Mangshan, Hunan). The distribution of the latter four species overlapped that of G. crenulata but are narrowly endemic, with only one locality known each. Our sampling covered 70% of the counties with the majority distributed across ten provinces/municipalities south of the Yangtze River in China. The number of individuals sampled from each population varied because of differences in population size and local occurrence frequency. Sample sizes ranged from six to 29 per site, with the exception of the EMS population (where only one individual was found) and the DWY population (where only two individuals were found). Despite their limited samples sizes, we deemed it important to include these populations in our analysis because they cover geographical areas not represented by other populations. The average sample size across all populations was 16 (Table 1). Within-population individuals were collected at least 30 m from one another to reduce spatial autocorrelation. Lu Lu undertook the formal identification of the plant material used in our study. Samples consisted of leaves dried with silica-gel desiccant. Specimen vouchers for each population were deposited in the Herbarium of the Kunming Institute of Botany, Chinese Academy of Sciences (KUN).

DNA amplification and sequencing

DNA was extracted from ca. 200 mg of each leaf sample with the cetyltrimethyl ammonium bromide protocol [80]. Two chloroplast DNA fragments (rpl33-psaJ and trnL-rpl32) and the two low-copy nuclear genes AAT and LOC (LCG) were employed (Table S8). PCR conditions were generally conducted as follows: 94 °C for four min, followed by 35 amplification cycles each at 94 °C for one min, annealing at 54–60 °C for 45 s, 72 °C for one min, and a final elongation at 72 °C for 10 min. The purified gene products were direct-sequenced with the Sanger method by using amplification primers and BigDye on an ABI 3730 capillary sequencer (Applied Biosystems, foster City, California, USA). Nuclear-gene PCR products that failed with direct sequencing were cloned into the pUC57-Kan vector (GenScript) for isolation of sequences. Eight single colonies per individual were selected randomly for colony PCR, and colonies containing target-size amplicons were sequenced with universal primers M13F and M13R (Sangon Biotech Co., Ltd., Shanghai, China). Sequences were edited and assembled with Sequencher [81] and Se-Al (University of Oxford, http://evolve.zoo.ox.ac.uk/software/seal/). All sequence data were deposited in GenBank (under accession numbers OR327752-OR333491).

Genetic diversity and differentiation statistics

Heterozygous sequences of the LCG were phased by using DnaSP 6.12 [82] with 1,000 runs, and a burn-in of 100, then analyzed with SPADS 1.0 [83] for calculating allelic richness (AR), nucleotide diversity (π), and interpopulation genetic differences (NST and GST). cpDNA sequences were analyzed with DnaSP 6.12 for haplotype diversity (Hd) and π, and with PERMUT 1.0 [84] for calculating genetic diversity (HT), intrapopulation genetic diversity (Hs), NST, and GST. Arlequin 3.11 [85] was used to calculate the coefficient of genetic differentiation (FST) among populations and estimate variation within populations.

Population genetic structure and migration

Based on our preliminary genetic analysis with STRUCTURE 2.3.4 [86], we divided the 89 population samples into the Western China (WC) group and the Eastern China (EC) group, separating the second-step and third-step regions. STRUCTURE was used to investigate the population structure of the LCG data based on an admixture model with correlated allele frequencies. The analysis involved 100,000 MCMC steps, with a burn-in of 50,000 and K values ranging from 1 to 10, repeated 20 times. We employed HARVST (http://taylor0.biology.ucla.edu/structureHarvester/) to calculate the posterior probability lnP (D) and its change rate ΔK to determine the best grouping [87]. To visualize the genetic similarity among populations, Principal coordinates analysis (PCoA) was performed with GenAlEx 6.54 [88].

For the network evolutionary relationships based on cpDNA data, we constructed haplotypes using the median linkage method in NETWORK 10.2 [89]. We also mapped the geographic distribution of the STRUCTURE grouping results and cpDNA haplotypes for each population using the ArcMap package in ArcGIS 10.2 (ESRI Inc, Redlands, California, USA). The Bayesian search strategy in MIGRATE 3.6.8 [90] was used with the LCG data to estimate potential historical gene flow (Nm) between the two groups. We calculated Θ = 4Nμ (N, effective population size; μ, mutation rate per generation) and M = m/μ (m, migration rate per generation) using the effective number of migrations. The parameters used were long-chains = 1, long-inc = 100, long-sample = 500,000, and burnin = 10,000. The initial values of Θ and M were derived from the FST values. We used SplitsTree 4.14.8 [91] to infer a neighbor net plot, visualizing the relationships among species from FST values of the nuclear and chloroplast data.

Population historical dynamics and divergence time estimates

We calculated Tajima’s D test [92] and Fu & Li’s D* test [93] to assess neutral evolution for both the LCG and cpDNA data. Mismatch distribution analysis was performed with Arlequin 3.11 by calculating Harpending’s raggedness index (HRag) and the sum of deviation squares (SSD) with 500 repetitions [94, 95] for either the overall data (populations) or for the two groups.

Colonization pathways based on the LCG data were analyzed by DIYABC Random Forest 1.0 (the DIYABC-RF analysis) [96]. Model construction was based on both taxonomic delimitation [22] and the geographical ranges of the populations (Fig. 1, (Fig. S1). The populations of the Gaultheria crenulata group were therefore classified into six groups: G. crenulata from the WC group (Pop1; denoted WC), G. crenulata from the EC group (Pop2; EC), G. wuliangshanensis (Pop3; JD), G. pingbienensis (Pop4; DWS), G. luchunensis (Pop5; YLC), and G. mangshanensis (Pop6; NMS). In addition, the results of population differentiation were used to pool genomically similar geographical sites and help guide the building of the models (Fig S3). We used the “random forest” method to classify the most likely scenario and estimate parameters, based on 20,000 and 50,000 simulations, respectively. We conducted 1,000 random forest trees and set uniform distributions as priors for all parameters. Generations time was based on the average five-year reproductive maturity period for Gaultheria [26] (Tables S4–S6).

Because of the low mutation rate of cpDNA data, the DIYABC-RF analysis could not be applied. We therefore used BEAST 2.4.2 [97] to estimate the haplotype divergence time for cpDNA differentiation. Gaultheria fragrantissima Wall. was selected as outgroup, and the calibrated Yule prior, strict clock model, and normal-distributed rate set were used. The program jModelTest 2.1.1 [98] yielded HKY + I + G model as the best model. Tracer 1.6 [99] was employed to validate the effective sample sizes (ESS > 200) for all prior and posterior distributions. As a secondary calibration point, we used the stem node age of the G. crenulata group [100] with a priori mean = 4.06 Ma, SD = 1 Ma, and 95% CI = 1.7–7.59 Ma on the fossil-calibrated phylogeny of the tribe Gaultherieae. The MCMC run consisted of 1 × 10 [7] generations, with sampling occurring every 1000 generations. The first 25% of samples were discarded as burn-in. Tree files were displayed with Figtree 1.4.3 (http://tree.bio.ed.ac.uk/software/figtree/).

IBD and IBE analyses

To investigate the influence of geography and/or environment on genetic differentiation, we tested isolation-by-distance (IBD) and isolation-by-environment (IBE) for the overall (populations) and for the WC/EC groups. For the LCG data, we assessed genetic distance using the values of pairwise differentiation (FST) and FST/ (1 − FST) calculated with Arlequin. For cpDNA data, we assessed genetic distance through FST and nucleotide divergence (DXY) using DnaSP. We characterized major environmental variations using bioclimatic variables obtained from WorldClim (http://worldclim.org) via Principal Component Analysis (PCA) with PC1 and PC2. We computed environmental and geographical distances using “vegdist” in the “vegan” package of R 4.05 (https://www.r-project.org/) and GenAlEx 6.54, respectively. To assess contributions, we used multiple matrix regression with randomization (MMRR) analysis to evaluate geographic distance and environmental variables [101]. We estimated correlations between genetic distance and effects from either geographic distance (with environmental variables removed) or environmental variables (with geographical distances removed) using partial Mantel tests. All analyses were conducted in R with 10,000 iterations.

Ecological niche modeling

Climate data of the Last Glacial Maximum (LGM; ca. 21 kya), mid-Holocene (MH; ca. 6 kya), current (1950–2000), and future (2070) were obtained from WorldClim, each period containing 19 bioclimatic variables with a 2.5 arc-minute spatial resolution. We simulated the LGM, MH, and current data using the CCSM4 model, whereas the future data was based on the RCP 8.5 scenario [102]. Considering the genetic structure may affect the results, we used GPS data of the G. crenulata group to generate species distribution models for the WC and EC groups, respectively. We identified correlated environmental variables using Pearson’s correlation coefficient algorithm, with a cutoff of |0.8| to prevent model overfitting [103]. Selected bioclimatic variables (bio1, bio2, bio5, bio6, bio7, bio8, bio12, bio14, and bio18) were imported into MaxEnt 3.3.3, with 75% of the data used for training and 25% for calibration; defaults were used for the other parameters. Model validation was replicate for 10 times independently, maximum iterations = 5000 with a convergence threshold 10−5. We evaluated each model using the area under the curve (AUC) of the receiver operating characteristic [104]. Maxent files were exported as a logistic output layer consisting of a grid map with ArcGIS.

Availability of data and materials

Data is provided within the manuscript or supplementary information files.

Abbreviations

A R :

Allelic richness

AUC:

The area under the curve

cpDNA:

Chloroplast DNA

EC:

Eastern China

ESS:

Effective sample sizes

F ST :

Genetic differentiation

Hd:

Haplotype diversity

H S :

Intrapopulation genetic diversity

H T :

Genetic diversity

IBD:

Isolation-by-distance

IBE:

Isolation-by-environment

LCG:

Low-copy nuclear genes

LGM:

Last Glacial Maximum

MH:

Mid-Holocene

Nm:

Gene flow

PCoA:

Principal coordinates analysis

SSD:

The sum of deviation squares

WC:

Western China

π :

Nucleotide diversity

References

  1. Wu ZY (Ed.). Vegetation of China. Science Press, Beijing. (In Chinese). 1980.

  2. Wang XH, Yan ER, Yan X, Wang LY. Analysis of degraded evergreen broad-leaved forest communities in Eastern China and issues in forest restoration. Acta Ecol Sin. 2005;25(7):1796–803.

    Google Scholar 

  3. Chen LZ, Sun H, Guo K. Floristic and Vegetational Geography of China. Beijing: Science Press; 2015.

    Google Scholar 

  4. Zhang XW, Li Y, Zhang Q, Fang YM. Ancient east-west divergence, recent admixture, and multiple marginal refugia shape genetic structure of a widespread oak species (Quercus acutissima) in China. Tree Genet Genomes. 2018;14(6):1–15.

    Article  Google Scholar 

  5. Qian H, Jin Y, Ricklefs RE. Phylogenetic diversity anomaly in angiosperms between eastern Asia and eastern North America. Proc Natl Acad Sci USA. 2017;114(43):11452–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Wang XH, Kent M, Fang XF. Evergreen broad-leaved forest in Eastern China: its ecology and conservation and the importance of resprouting in forest restoration. For Ecol Manag. 2007;245(1–3):76–87.

    Article  Google Scholar 

  7. Feng G, Sevnning JC, Mi XC, Jia Q, Rao MD, Ren HB, Bebber DP, Ma KP. Anthropogenic disturbance shapes phylogenetic and functional tree community structure in a subtropical forest. For Ecol Manag. 2014;313(2):188–98.

    Article  Google Scholar 

  8. Wu A, Deng X, He H, Ren X, Jing Y, Xiang W, Ouyang S, Yan WD, Fang X. Responses of species abundance distribution patterns to spatial scaling in subtropical secondary forests. Ecol Evol. 2019;9(9):5338–47.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Xu WB, Svenning JC, Chen GK, Zhang MG, Huang JH, Chen B, Ordonez A, Ma KP. Human activities have opposing effects on distributions of narrow-ranged and widespread plant species in China. Proc Natl Acad Sci USA. 2019;116(52):26674–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Zuo Z, Zheng X. Forest structure and tree species diversity across a disturbance gradient in evergreen broadleaved secondary forests. Journal of Zhejiang A&F University. 2019;36(1):21–30.

    Google Scholar 

  11. Norden N, Chazdon RL, Chao A, Jiang YH, Vílchez-Alvarado B. Resilience of tropical rain forests: tree community reassembly in secondary forests. Ecol Lett. 2009;12(5):385–94.

    Article  PubMed  Google Scholar 

  12. Liu W, Xie J, Zhou H, Kong H, Hao G, Fritsch PW, Gong W. Population dynamics linked to glacial cycles in Cercis chuniana FP Metcalf (Fabaceae) endemic to the montane regions of subtropical China. Evol Appl. 2021;14(11):2647–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Wehenkel C, Corral-Rivas JJ, Hernandez-Diaz JC. Genetic diversity in relation to secondary succession of forest tree communities. Pol J Ecol. 2011;59(1):45–54.

    Google Scholar 

  14. Lian C, Oishi R, Miyashita N, Nara K, Nakaya H, Wu B, Zhou Z, Hogetsu T. Genetic structure and reproduction dynamics of Salix reinii during primary succession on Mount Fuji, as revealed by nuclear and chloroplast microsatellite analysis. Mol Ecol. 2003;12(3):609–18.

    Article  CAS  PubMed  Google Scholar 

  15. Chen XD, Yang J, Guo YF, Zhao YM, Zhou T, Zhang X, Zhao GF. Spatial genetic structure and demographic history of the dominant forest oak Quercus fabri hance in subtropical China. Front Plant Sci. 2021;11: 583284.

    Article  PubMed  PubMed Central  Google Scholar 

  16. Sun R, Lin F, Huang P, Ye X, Lai J, Zheng Y. Phylogeographical structure of Liquidambar formosana Hance revealed by chloroplast phylogeography and species distribution models. Forests. 2019;10(10):858.

    Article  Google Scholar 

  17. Ricklefs RE, Latham RE. Intercontinental correlation of geographical ranges suggests stasis in ecological traits of relict genera of temperate perennial herbs. Am Nat. 1992;139(6):1305–21.

    Article  Google Scholar 

  18. Luo Y, Hu H, Zhao M, Li H, Liu S, Fang J. Latitudinal pattern and the driving factors of leaf functional traits in 185 shrub species across eastern China. J Plant Ecol. 2019;12(1):67–77.

    Article  Google Scholar 

  19. Li Y, Ni Y, Xu H, Lian J, Ye W. Relationship between variation of plant functional traits and individual growth at different vertical layers in a subtropical evergreen broad-leaved forest of Dinghushan. Biodiv Sci. 2021;29(9):1186.

    Article  Google Scholar 

  20. Myers-Smith IH, Elmendorf SC, Beck PS, Wilmking M, Hallinger M, Blok D, Tape KD, Rayback SA, Macias-Fauria M, Forbes BC, Speed JDM, Boulanger-Lapointe N, Rixen C, Lévesque E, Schmidt NM, Baittinger C, Trant AJ, Hermanutz L, Collier LS, Dawes MA, Lantz TC, Weijers S, Jørgensen RH, Buchwal A, Buras A, Naito AT, Ravolainen V, Schaepman-Strub G, Wheeler JA, Wipf S, Guay KC, Hik DS, Vellend M. Climate sensitivity of shrub growth across the tundra biome. Nat Clim Change. 2015;5(9):887–91.

    Article  Google Scholar 

  21. Ropars P, Angers‐Blondin S, Gagnon M, Myers‐Smith IH, Lévesque E, Boudreau S. Different parts, different stories: climate sensitivity of growth is stronger in root collars vs. stems in tundra shrubs. Glob Change Biol. 2017; 23(8): 3281–3291.

  22. Li YR, Xu YL, Fritsch PW, Lu L. Patterns of genetic variation and morphology support the recognition of five species in the Gaultheria leucocarpa Blume (Ericaceae) group from mainland China. Ecol Evol. 2023;13: e10178.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Nikolić M, Marković T, Mojović M, Pejin B, Savić A, Perić T, Marković D, Stević T, Soković M. Chemical composition and biological activity of Gaultheria procumbens L. essential oil. Ind Crop Prod. 2013; 49: 561–567.

  24. Liu WR, Qiao WL, Liu ZZ, Wang XH, Jiang R, Li SY. Gaultheria: Phytochemical and pharmacological characteristics. Molecules. 2013;18(10):12071–108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Lu L, Fritsch PW, Cruz BC, Wang H, Li DZ. Reticulate evolution, cryptic species, and character convergence in the core East Asian clade of Gaultheria (Ericaceae). Mol Phylogenet Evol. 2010;57:364–79.

    Article  PubMed  Google Scholar 

  26. Bunnell FL. Reproduction of salal (Gaultheria shallon) under forest canopy. Can J For Res. 1990;20(1):91–100.

    Article  Google Scholar 

  27. Wang L, Li Y, Noshiro S, Suzuki M, Arai T, Kobayashi K, Xie L, Zhang MY, He N, Fang YM, Zhang F. Stepped geomorphology shaped the phylogeographic structure of a widespread tree species (Toxicodendron vernicifluum, Anacardiaceae) in East Asia. Front Plant Sci. 2022;13:1–17.

    Google Scholar 

  28. Després L. One, two or more species? Mitonuclear discordance and species delimitation. Mol Ecol. 2019;28:3845–7.

    Article  PubMed  Google Scholar 

  29. Hou H, Ye H, Wang Z, Wu J, Gao Y, Han W, Wang Y. Demographic history and genetic differentiation of an endemic and endangered Ulmus lamellosa (Ulmus). BMC Plant Biol. 2020;20(1):1–14.

    Article  CAS  Google Scholar 

  30. Chen XD, Yang J, Feng LI, Zhou T, Zhang H, Li HM, Zhao GF. Phylogeography and population dynamics of an endemic oak (Quercus fabri Hance) in subtropical China revealed by molecular data and ecological niche modeling. Tree Genet Genomes. 2020;16(1):1–13.

    Article  Google Scholar 

  31. Petit RJ, Duminil J, Fineschi S, Hampe A, Salvini D, Vendramin GG. Comparative organization of chloroplast, mitochondrial and nuclear diversity in plant populations. Mol Ecol. 2005;14:689–701.

    Article  CAS  PubMed  Google Scholar 

  32. Tanaka T. Species Problem in Citrus. Tokyo: Japanese Society for the promotion of science; 1954.

    Google Scholar 

  33. Clark MK, Schoenbohm LM, Royden LH, Whipple KX, Burchfiel BC, Zhang X, Chen L. Surface uplift, tectonics, and erosion of eastern Tibet from large-scale drainage patterns. Tectonics. 2004;23(1):1–20.

    Article  CAS  Google Scholar 

  34. Peng Z, Ho SY, Zhang Y, He S. Uplift of the Tibetan plateau: evidence from divergence times of glyptosternoid catfishes. Mol Phylogenet Evol. 2006;39(2):568–72.

    Article  CAS  PubMed  Google Scholar 

  35. Ming Q, Shi Z, Dong M. The inquiry on genesis and formation times of the first bend of Yangtze River. Prog Geogr. 2007;26(3):119–26.

    Google Scholar 

  36. Pan T, Wu S, He D, Dai E, Liu Y. Effects of longitudinal range-gorge terrain on the eco-geographical pattern in Southwest China. J Geogr Sci. 2012;22(5):825–42.

    Article  Google Scholar 

  37. Zhang TC, Comes HP, Sun H. Chloroplast phylogeography of Terminalia franchetii (Combretaceae) from the eastern Sino-Himalayan region and its correlation with historical river capture events. Mol Phylogenet Evol. 2011;60(1):1–12.

    Article  CAS  PubMed  Google Scholar 

  38. Rana SK, Luo D, Rana HK, O’Neill AR, Sun H. Geoclimatic factors influence the population genetic connectivity of Incarvillea arguta (Bignoniaceae) in the Himalaya-Hengduan Mountains biodiversity hotspot. J Syst Evol. 2021;59(1):151–68.

    Article  Google Scholar 

  39. Wang X, Tong L, Deng J, Li L, Xiang P, Xu L, Song Z. Insights into historical drainage evolution based on the phylogeography of Schizopygopsis malacanthus Herzenstein (Cypriniformes, Cyprinidae) across the upper and middle Yalong River drainage in the Hengduan Mountains region, southwest China. Glob Ecol Conserv. 2022;35: e02084.

    Google Scholar 

  40. Shafer AB, Wolf JB. Widespread evidence for incipient ecological speciation: a meta-analysis of isolation-by-ecology. Ecol Lett. 2013;16(7):940–50.

    Article  PubMed  Google Scholar 

  41. Sexton JP, Hangartner SB, Hoffmann AA. Genetic isolation by environment or distance: which pattern of gene flow is most common? Evolution. 2014;68(1):1–15.

    Article  CAS  PubMed  Google Scholar 

  42. Still DW, Kim DH, Aoyama N. Genetic variation in Echinacea angustifolia along a climatic gradient. Ann Bot. 2005;96(3):467–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Twyford AD, Wong ELY, Friedman J. Multi-level patterns of genetic structure and isolation by distance in the widespread plant Mimulus guttatus. Heredity. 2020;125(4):227–39.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Vargas-Mendoza CF, Ortegón-Campos I, Marrufo-Zapata D, Herrera CM, Parra-Tabla V. Genetic diversity, outcrossing rate, and demographic history along a climatic gradient in the ruderal plant Ruellia nudiflora (Acanthaceae). Rev Mex Biodivers. 2015;86(2):508–20.

    Article  Google Scholar 

  45. Hutchison DW, Templeton AR. Correlation of pairwise genetic and geographic distance measures: inferring the relative influences of gene flow and drift on the distribution of genetic variability. Evolution. 1999;53(6):1898–914.

    Article  PubMed  Google Scholar 

  46. Pertoldi C, Ruiz-Gonzalez A, Bahrndorff S, Lauridsen NR, Henriksen TN, Eskildsen A, Høye TT. Strong isolation by distance among local populations of an endangered butterfly species (Euphydryas aurinia). Ecol Evol. 2021;11(18):12790–800.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Shi MM, Michalski SG, Chen XY, Durka W. Isolation by elevation: genetic structure at neutral and putatively non-neutral loci in a dominant tree of subtropical forests, Castanopsis eyrei. PLoS ONE. 2011;6(6): e21302.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kaufman DS, Broadman E. Revisiting the Holocene global temperature conundrum. Nature. 2023;614(7948):425–35.

    Article  CAS  PubMed  Google Scholar 

  49. Zheng Z, Chen C, Huang K, Zhang X, Kershaw P, Cheng J, Li J, Yue YF, Wan QC, Zhang YZ, Tang YJ, Wang MY, Xiao XY, Cheddadi R. Holocene warming and evergreen/deciduous forest replacement across eastern China. Quat Sci Rev. 2023;307: 108057.

    Article  Google Scholar 

  50. Yin QY, Fan Q, Li P, Truong D, Zhao WY, Zhou RC, Chen SF, Liao WB. Neogene and Quaternary climate changes shaped the lineage differentiation and demographic history of Fokienia hodginsii (Cupressaceae s.l.), a Tertiary relict in East Asia. J Syst Evol. 2021; 59(5): 1081–1099.

  51. Kou Y, Fan D, Cheng S, Yang Y, Wang M, Wang Y, Zhang Z. Peripatric speciation within Torreya fargesii (Taxaceae) in the Hengduan Mountains inferred from multi-loci phylogeography. BMC Ecol Evol. 2023;23(1):74.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Cheng Y, Liu H, Wang H, Hao Q. Differentiated climate-driven Holocene biome migration in western and eastern China as mediated by topography. Earth Sci Rev. 2018;182:174–85.

    Article  Google Scholar 

  53. Jacobsen D. The dilemma of altitudinal shifts: caught between high temperature and low oxygen. Front Ecol Environ. 2020;18(4):211–8.

    Article  Google Scholar 

  54. Seppä P, Johansson H, Gyllenstrand N, Pálsson S, Pamilo P. Mosaic structure of native ant supercolonies. Mol Ecol. 2012;21(23):5880–91.

    Article  PubMed  Google Scholar 

  55. Bai WN, Wang WT, Zhang DY. Contrasts between the phylogeographic patterns of chloroplast and nuclear DNA highlight a role for pollen-mediated gene flow in preventing population divergence in an East Asian temperate tree. Mol Phylogenet Evol. 2014;81:37–48.

    Article  PubMed  Google Scholar 

  56. Fu MJ, Wu HY, Jia DR, Tian B. Evolutionary history of a desert perennial Arnebia szechenyi (Boraginaceae): Intraspecific divergence, regional expansion and asymmetric gene flow. Plant Divers. 2021;43(6):462–71.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Gao J, Liu ZL, Zhao W, Tomlinson KW, Xia SW, Zeng QY, Wang XR, Chen J. Combined genotype and phenotype analyses reveal patterns of genomic adaptation to local environments in the subtropical oak Quercus acutissima. J Syst Evol. 2021;59(3):541–56.

    Article  Google Scholar 

  58. Mirick S, Quinn JA. Some observations on the reproductive biology of Gaultheria procumbens (Ericaceae). Am J Bot. 1981;68(10):1298–305.

    Article  Google Scholar 

  59. Primack RB. Insect pollination in the New Zealand Mountain flora. N Z J Bot. 1983;21(3):317–33.

    Article  Google Scholar 

  60. Kong H, Condamine FL, Harris AJ, Chen J, Pan B, Möller M, Kang M. Both temperature fluctuations and East Asian monsoons have driven plant diversification in the karst ecosystems from southern China. Mol Ecol. 2017;26(22):6414–29.

    Article  PubMed  Google Scholar 

  61. Li Y, Zhang X, Fang Y. Landscape features and climatic forces shape the genetic structure and evolutionary history of an oak species (Quercus chenii) in East China. Front Plant Sci. 2019;10:1060.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Fu J, Wen L. Impacts of Quaternary glaciation, geological history and geography on animal species history in continental East Asia: a phylogeographic review. Mol Ecol. 2023;32:4497–514.

    Article  PubMed  Google Scholar 

  63. Zhang H, Lu J, Tang S, Huang Z, Cui L, Lan D, Wang H, Hou R, Xiao W, Guo ST, He G, Huang K, Zhang P, Pan H, Oxnard C, Pan RL, Li B. Southwest China, the last refuge of continental primates in East Asia. Biol Conserv. 2022;273: 109681.

    Article  Google Scholar 

  64. Brunetti M, Magoga G, Iannella M, Biondi M, Montagna M. Phylogeography and species distribution modelling of Cryptocephalusbarii (Coleoptera: Chrysomelidae): is this alpine endemic species close to extinction? ZooKeys. 2019;856:3.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Ao H, Ruan J, Martinón-Torres M, Krapp M, Liebrand D, Dekkers MJ, Caley T, Jonell TN, Zhu Z, C Huang, Li X, Zhang Z, Sun Q, Yang P, Jiang J, Li X, Xie X, Song Y, Qiang X, Zhang P, An Z. Concurrent Asian monsoon strengthening and early modern human dispersal to East Asia during the last interglacial. P Proc Natl Acad Sci USA. 2024; 121(3): e2308994121.

  66. Zhisheng A, Kutzbach JE, Prell WL, Porter SC. Evolution of Asian monsoons and phased uplift of the Himalaya-Tibetan plateau since Late Miocene times. Nature. 2001;411:62–6.

    Article  CAS  PubMed  Google Scholar 

  67. Yang A, Zhong Y, Liu S, Liu L, Liu T, Li Y, Yu F. 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. 2019;7: e6355.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Cheng S, Zeng W, Fan D, Liang H, Yang Y, Kou Y, Zhang Z. Subtle east–west phylogeographic break of Asteropyrum (Ranunculaceae) in subtropical China and adjacent areas. Diversity. 2021;13(12):627.

    Article  CAS  Google Scholar 

  69. Shen Y, Cheng Y, Li K, Li H. Integrating phylogeographic analysis and geospatial methods to infer historical dispersal routes and glacial refugia of Liriodendron chinense. Forests. 2019;10(7):565.

    Article  Google Scholar 

  70. Huang X, Chen J, Yang X, Duan S, Long C, Ge G, Rong J. Low genetic differentiation among altitudes in wild Camellia oleifera, a subtropical evergreen hexaploid plant. Tree Genet Genomes. 2018;14:1–12.

    Article  Google Scholar 

  71. He X, Ziegler AD, Elsen PR, Feng Y, Baker JC, Liang S, Holden J, Spracklen DV, Zeng Z. Accelerating global mountain forest loss threatens biodiversity hotspots. One Earth. 2023;6(3):303–15.

    Article  Google Scholar 

  72. Pinto AV, Hansson B, Patramanis I, Morales HE, Oosterhout C. The impact of habitat loss and population fragmentation on genomic erosion. Conserv Genet. 2024;25:49–57.

    Article  CAS  Google Scholar 

  73. Silveira dos Santos J, Vitorino LC, Fabrega Gonçalves R, Corrêa Côrtes M, Souza Cruz Alves R, Cezar Ribeiro M, Garcia Collevatti R. Matrix dominance and landscape resistance affect genetic variability and differentiation of an Atlantic Forest pioneer tree. Landsc Ecol. 2022; 37(10): 2481–2501.

  74. Qin SY, Zuo ZY, Xu SX, Liu J, Yang FM, Luo YH, Ye JW, Zhao Y, Rong J, Liu B, Ma PF, Li DZ. Anthropogenic disturbance driving population decline of a dominant tree in East Asia evergreen broadleaved forests over the last 11,000 years. Conserv Biol. 2023; e14180.

  75. Exposito-Alonso M, Booker TR, Czech L, Gillespie L, Hateley S, Kyriazis CC, Lang PLM, Leventhal L, Nogues-Bravo D, Pagowski V, Ruffley M, Spence JP, Arana SET, Weiß CL, Zess E. Genetic diversity loss in the Anthropocene. Science. 2022;377(6613):1431–5.

    Article  CAS  PubMed  Google Scholar 

  76. Jia Y, Milne RI, Zhu J, Gao LM, Zhu GF, Zhao GF, Liu J, Li ZH. Evolutionary legacy of a forest plantation tree species (Pinus armandii): Implications for widespread afforestation. Evol Appl. 2020;13(10):2646–62.

    Article  PubMed  PubMed Central  Google Scholar 

  77. Ma ZF, Pen J, Gao WL, Tian H. Climate Variation of Southwest China in Recent 40 Years. Plateau Meteorol. 2006;25(4):79–88.

    Google Scholar 

  78. Ma XJ, Zheng JH, Chen XZ. Studies on Resources of Ethnomedicine Gaultheria leucocarpa var. yunnanensis. China J Chin Mater Med. 2001;26(2):85–9.

    CAS  Google Scholar 

  79. Fenster CB, Ballou JD, Dudash MR, Eldridge MDB, Frankham R, Lacy RC, Ralls K, Sunnucks P. Conservation and genetics. Yale J Biol Med. 2018;91(4):491–501.

    PubMed  PubMed Central  Google Scholar 

  80. Doyle JJ, Doyle JL. A rapid DNA isolation procedure for small quantities of fresh leaf material. Phytochem Bull. 1987;19:11–5.

    Google Scholar 

  81. Nishimura D. Sequencher 3.1.1. Biotech Softw Internet Rep. 2004; 1: 1–2.

  82. Rozas J, Ferrer-Mata A, Sánchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE, Sánchez-Gracia A. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol Biol Evol. 2017;34(12):3299–302.

    Article  CAS  PubMed  Google Scholar 

  83. Dellicour S, Mardulyn P. SPADS 1.0: a toolbox to perform spatial analyses on DNA sequence data sets. Mol Ecol Resour. 2014;14(3):647–51.

    Article  PubMed  Google Scholar 

  84. Pons O, Petit RJ. Measuring and testing genetic differentiation with ordered versus unordered alleles. Genetics. 1996;144(3):1237–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Excoffier L, Laval G, Schneider S. Arlequin (version 3.0): an integrated software package for population genetics data analysis. Evol Bioinform. 2005;1:47–50.

  86. Hubisz M. Inferring weak population structure with the assistance of sample group information. Mol Ecol Resour. 2010;9(5):1322–32.

    Article  Google Scholar 

  87. Earl DA, Vonholdt BM. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genet Resour. 2012;4:359–61.

    Article  Google Scholar 

  88. Peakall ROD, Smouse PE. GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Mol Ecol Notes. 2006;6(1):288–95.

    Article  Google Scholar 

  89. Forster P, Forster L, Renfrew C, Forster M. Phylogenetic network analysis of SARS-CoV-2 genomes. Proc Natl Acad Sci USA. 2020;117(17):9241–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Beerli P. Comparison of Bayesian and maximum likelihood inference of population genetic parameters. Bioinformatics. 2006;22(3):341–5.

    Article  CAS  PubMed  Google Scholar 

  91. Huson DH, Auch AF, Qi J, Schuster SC. MEGAN analysis of metagenomic data. Genome Res. 2007;17(3):377–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Tajima F. Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics. 1989;123(3):585–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Fu YX, Li WH. Statistical tests of neutrality of mutations. Genetics. 1993;133(3):693–709.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Rogers AR, Harpending H. Population growth makes waves in the distribution of pairwise genetic differences. Mol Biol Evol. 1992;9(3):552–69.

    CAS  PubMed  Google Scholar 

  95. Harpending HC. Signature of ancient population growth in a low-resolution mitochondrial DNA mismatch distribution. Hum Biol. 1994;6(4):591–600.

    Google Scholar 

  96. Collin FD, Durif G, Raynal L, Lombaert E, Gautier M, Vitalis R, Estoup A. Extending approximate Bayesian computation with supervised machine learning to infer demographic history from genetic polymorphisms using DIYABC Random Forest. Mol Ecol Resour. 2021;21(8):2598–613.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Bouckaert R, Heled J, Kühnert D, Vaughan T, Wu CH, Xie D, Drummond AJ. BEAST 2: a software platform for Bayesian evolutionary analysis. PLoS Comput Biol. 2014;10(4): e1003537.

    Article  PubMed  PubMed Central  Google Scholar 

  98. Darriba D, Taboada GL, Doallo R, Posada D. jModelTest 2: more models, new heuristics and parallel computing. Nat Methods. 2012;9(8):772–772.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Rambaut A, Drummond AJ, Xie D, Baele G, Suchard MA. Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Syst Biol. 2018;67(5):901–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Lu L, Fritsch PW, Matzke NJ, Wang H, Kron KA, Li DZ. Why is fruit colour so variable? Phylogenetic analyses reveal relationships between fruit-colour evolution, biogeography and diversification. Global Ecol Biogeogr. 2019;28:891–903.

    Article  Google Scholar 

  101. Wang IJ. Examining the full effects of landscape heterogeneity on spatial genetic variation: a multiple matrix regression approach for quantifying geographic and ecological isolation. Evolution. 2013;67(12):3403–11.

    Article  PubMed  Google Scholar 

  102. Gent PR, Danabasoglu G, Donner LJ, Holland MM, Hunke EC, Jayne SR, Zhang M. The community climate system model version 4. J Clim. 2011;24(19):4973–91.

    Article  Google Scholar 

  103. Phillips SJ, Anderson RP, Schapire RE. Maximum entropy modeling of species geographic distributions. Ecol Modell. 2006;190(3–4):231–59.

    Article  Google Scholar 

  104. Fawcett T. An introduction to ROC analysis. Pattern Recognit Lett. 2006;27(8):861–74.

    Article  Google Scholar 

Download references

Acknowledgements

The authors express their gratitude to Yan-Quan Chen and Guo-Hong Li for their assistance in sample collection, in particular to Mr. Jun Chen for collecting the novel species Gaultheria mangshanensis in Mangshan National Nature Reserve.

Funding

This study was supported by grants from National Natural Science Foundation China (31960080, 41671052, and 42175139), Program Innovative Research Team in Science and Technology in Yunnan Province (No. 202005AE160004), Reserve talents of young and middle-aged academic and technical leaders in Yunnan Province (No. 202005AC160020), and the U.S. National Science Foundation (DEB-0717711 to PWF).

Author information

Authors and Affiliations

Authors

Contributions

LL and ZLD designed the research project; LL, YRL, and XJC performed field collections; YRL, ZLD and GGZ generated molecular data and performed data analyses; YRL, LL, and PWF wrote the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Zhao-Li Ding or Lu Lu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, YR., Fritsch, P.W., Zhao, GG. et al. Population differentiation and dynamics of five pioneer species of Gaultheria from the secondary forests in subtropical China. BMC Plant Biol 24, 516 (2024). https://doi.org/10.1186/s12870-024-05189-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-024-05189-z

Keywords