Open Access
How to translate text using browser tools
11 March 2022 Genetic Diversity and Population Structure of Rhododendron longipedicellatum, an Endangered Species
Yurong Cao, Yongpeng Ma, Zhenghong Li, Xiongfang Liu, Detuan Liu, Suping Qu, Hong Ma
Author Affiliations +
Abstract

Rhododendron longipedicellatum is an endangered species endemic to southeastern Yunnan, China. Assessment of genetic variation is critical for protecting endangered species. Therefore, we used EST-SSR markers to analyze the genetic characteristics of R. longipedicellatum. The results revealed high genetic diversity at the species level (He = 0.559, NA = 9.529) and within populations (He = 0.507, NA = 5.910) and moderate genetic differentiation between populations (FST = 0.083). In addition, more genetic variation existed within populations (91.25%) compared with variation among populations (8.75%). The STRUCTURE analysis showed that 150 individuals from five existing populations could best be divided into two genetic groups. At the population level, the neighbor-joining (NJ) tree and principal coordinates analysis (PCoA) analyses also divided them into two groups. In addition, Bottleneck analyses using the Two-Phase Model (TPM) and Stepwise Mutation Model (SMM) as well as the Garza-Williamson Index revealed widespread signatures of bottleneck events. These results provide vital information for scientifically formulating conservation strategies for the endangered R. longipedicellatum.

Introduction

Genetic diversity is a key component of biodiversity. The study of genetic diversity and population structure of species is the theoretical basis for scientific and rational utilization of germplasm resources and improvement of genetic characteristics, as well as the prerequisite for effective conservation measures (Frankham, 2003; Mable, 2019). With the rapid development and continuous improvement of techniques in molecular biology, molecular markers have been widely used for population genetic diversity analysis, breeding programs, and evolutionary research (Ma et al., 2021; Yadav et al., 2019). Among the molecular marker types, simple sequence repeat (SSR) markers, including genomic SSR (gSSR) and expressed sequence tag SSR (EST-SSR), are highly regarded because of their rich content distribution, high polymorphism, ease of use, and repeatability (Liu et al., 2019; Wei et al., 2011). Compared with gSSR, EST-SSR have a higher level of transferability to related species, are more evolutionarily conserved, have lower development costs, and closer relationships with genes of known functions (Singh et al., 2013; Zhang et al., 2017).

Rhododendron is the largest genus in Ericaceae, with the widest distribution in the Northern Hemisphere, consisting of eight subgenera and more than 1025 species. China has about 581 species from six subgenera, of which 421 species are endemic to China (Cai et al., 2016; Ma et al., 2017). China is the most important modern distribution and evolution center of Rhododendron species, but the loss and destruction of wild resources has become a serious threat in recent years, and many endemic and rare endangered species are extinct or on the verge of extinction (Ma et al., 2014). Although much research has been conducted on the genetic structure and population dynamics of Rhododendron, few studies have been conducted on endangered species whose habitats are seriously threatened. Thus, conservation research on endemic and endangered Rhododendron species is urgently needed.

Rhododendron longipedicellatum is endemic to Southeast Yunnan, China (Cai et al., 2016). Its beautiful tree shape, bright pure yellow flowers, and unusual flowering time give this species a high value for development and utilization in landscaping. Since 2014, our study group has found only five wild populations with less than 1000 plants remaining in limestone habitats in Malipo County, Yunnan Province at an altitude of approximately 1183–1316 m. Due to the proximity to human settlements and high ornamental value, additional habitat loss and removal of wild plants is a relatively serious threat to conserving these remaining populations. Furthermore, within the narrow geographic distribution, the number of populations and individual plants is small, few true seedlings are present within each population, and natural regeneration is weak. At present, the wild, undeveloped habitat available to sustain populations such as ZWL and XL is continuously decreasing ( Supplementary Figure S1). According to the IUCN Red List Categories and Criteria (IUCN, 2001), R. longipedicellatum is already extremely endangered (CR B1ab [i, iii, v]) and requires urgent protection. Thus, this study used 17 previously developed high-polymorphism EST-SSR markers (Li, Liu, Li, Wan, et al., 2018) to study the genetic characteristics of 150 individuals from five residual populations of R. longipedicellatum, attempting to assess the following: (1) genetic diversity and structure; (2) gene flow between populations; (3) correlation between genetic distance and geographical distance; and (4) bottleneck effects. Findings are expected to illustrate the possible genetic consequences of residual small populations in the limestone habitat so as to provide a reference for effective conservation and germplasm resource management of R. longipedicellatum.

Methods

Experimental Materials

Young leaves of 150 healthy mature plants (30 samples from each population) were collected from five residual populations. Locality and voucher information for samples are provided in  Supplementary Table S1 and Figure 1. Samples were collected at a distance of at least 2.5 m to avoid collecting clone-mates. The collected leaf tissues were quickly dried with silica gel and stored at −20°C until further use.

Figure 1.

R. longipedicellatum (A) and the geographical distribution of the five extant populations of R. longipedicellatum (B). Note: ZWL, XL, WJL, XCW, and WBL are population codes.

10.1177_19400829221078112-fig1.tif

DNA Extraction and EST-SSR Genotyping

A modified cetyltrimethylammonium bromide (CTAB) method was used to extract genomic DNA (Doyle & Doyle, 1987). In brief, gel-dried fragments of leaves (0.1 g) were placed in liquid nitrogen, quickly ground to a powder, and then mixed with 1 mL of CTAB buffer (100 mm Tris-HCl, pH 8.0, 20 mm EDTA, 2% CTAB) with 0.1% β-mercaptoethanol added immediately prior to use. The mixture was incubated at 65°C for 60 min and centrifuged for 2 min at 13,000 rpm, after which 600 µL of the supernatant was transferred to a 1.5 mL microcentrifuge tube. The supernatant was mixed twice with phenol/chloroform/isoamyl alcohol (25:24:1) and mixed for 10 min, followed by centrifugation for 10 min at 12,000 rpm. DNA was precipitated from the aqueous phase by adding 0.1 mL of 3 m sodium acetate and 2.5 mL of ethanol. The DNA precipitate was washed twice with 70% ethanol and dissolved in water. DNA quantity was assessed using 0.8% agarose gel electrophoresis ( Supplementary Figure S2).

All samples were genotyped using 17 EST-SSR markers previously developed for this species ( Supplementary Table S1) (Li, Liu, Li, Wan, et al., 2018). For all markers, PCR amplification (three replicates of each sample) was performed on an ABI 3730 Thermal Cycler (Thermo Fisher Scientific, Waltham, Massachusetts, USA). PCR amplifications were performed in a final volume of 10 µL, containing 1 µL (10–30 ng) of template DNA, 5 µL of 0.7× Multiplex PCR Master Mix (QIAGEN, Hilden, Germany), 0.5 µL (10 pM) of each primer, and 3 µL of RNase-free water. The PCR program was as follows: 5 min at 94°C; 30 cycles of 30 s at 94°C, annealing at 58°C for 30 s, and 30 s at 72°C; and a final extension step at 72°C for 5 min. The PCR products were pooled and purified using the Wizard SV Gel and PCR Clean-Up System (Promega), following the manufacturer’s instructions. The 5′ end of each forward primer for the 17 markers was tagged with one of three fluorescent dyes (FAM, HEX, or ROX), and multiplex PCR amplifications were performed for the individuals of R. longipedicellatum using the PCR conditions described above. The fluorescence signal detected by capillary electrophoresis was interpreted using the GeneMapper v4.1. All individual-specific bands corresponding to each primer were read out to establish the original data matrix. Data editing and format conversion were carried out in Convert v1.31 and Genalex v6.2.

Data Analysis

The frequency of null alleles was estimated for each locus in each population using INEST 2.2 (Chybicki & Burczyk, 2009). Genepop v4.2 was used to calculate Linkage disequilibrium (LD) for all locus pairs and Hardy-Weinberg equilibrium (HWE) in each population, with deviations from expected proportions evaluated with a Bonferroni corrected p-value threshold of 0.05 (Rice, 1989). The genetic diversity indices were calculated in PopGene v1.31 (Yeh et al., 1999), including the average number of alleles (NA), the number of effective alleles (NE), Shannon information index (I), observed heterozygosity (Ho), expected heterozygosity (He), inbreeding coefficient (FIS).

We adopted the Markov Chain Monte Carlo (MCMC) method and used STRUCTURE v2.3.4 (Pritchard et al., 2000) to evaluate genetic structure and shared ancestry across the five putative population groups. The analysis was performed using the admixture model of ancestry, which assumes independent origins of each allele, with operating parameter settings “Length of Burnin period” and “Number of MCMC Reps after Burnin” in the running length set to 30,000 and 100,000, respectively. The K value range was set at 1–20, and 20 independent runs were conducted for each K value. The most likely number of clusters was evaluated using both ΔK and the log-likelihood value with STRUCTURE HARVESTER v0.6.94 (Earl & VonHoldt, 2012; Evanno et al., 2005). Following this, cluster analysis based on shared allele distance (DAS) was performed using Population v1.2.30 (Langella, 2007) with the neighbor-joining (NJ) method with 1000 bootstraps. The clustering tree was edited using the FigTree v1.4.0. Based on Nei’s genetic distance (Nei, 1973), principal coordinates analysis (PCoA) of populations was conducted with ape package v5.4 (Paradis & Schliep, 2019). The genetic distance and geographic distance matrices between populations were computed using PopGene v1.31 and Geographic Distance Matrix Generator v1.2.2, respectively, and the Mantel test was carried out with PASSaGE v2 (999 permutations were calculated).

Analysis of molecular variance (AMOVA) was carried out to estimate the partitioning of genetic variation between and within the populations using Arlequin v3.5 (Excoffier & Lischer, 2010). The genetic differentiation between populations was evaluated by calculating the genetic differentiation index (FST; Weir & Cockerham, 1984) using FSTAT v.2.9.3 (Goudet, 2001). Gene flow (Nm) was calculated by the formula Nm = ([1/Fst] − 1)/4 (Slatkin, 1985). The Garza-Williamson index (G-W) was calculated using Arlequin v3.5 (Excoffier & Lischer, 2010). The data were assessed for a population genetic bottleneck signature using the Sign test and Wilcoxon test under the Stepwise Mutation Model (SMM) and Two-Phase Model (TPM) in Bottleneck v1.2 (Cornuet & Luikart, 1997; Piry et al., 1999). The variance for TPM was set to 30, and the proportion of SMM in TPM was 70%, with 1000 iterations. The remaining relevant parameters were set as the software default values.

Results

Genetic Diversity

Using INEST, the frequency of null alleles was estimated to be lower than the threshold of 0.05 at each of the 17 markers across populations (Table 1) (Dakin & Avise, 2004; Y. Li et al., 2019). No consistent genotypic disequilibrium was found between any two loci across all populations; therefore, all loci were used for further analyses. Significant deviation from HWE was only detected in one of the five populations (p < 0.05, after Bonferroni correction; Table 1). At the population level, the five R. longipedicellatum populations showed high genetic diversity (Table 1). The average number of alleles per locus was high (9.529) across all populations, while individual population values ranged from 4.647 to 6.177. The average number of alleles was higher in the pooled population data set because unique alleles were present in some of the populations. Shannon’s information index as a measure of gene diversity was similar across populations, ranging from 1.207 to 0.860. In general, the population XCW contained the greatest level of genetic diversity, and WBL the least.

Table 1.

Genetic Diversity Indices for Five Populations of R. longipedicellatum.

Genetic Structure

The results of the AMOVA analysis of the five populations are presented in Table 2. Of the total genetic variation, 91.25% occurred within populations, and only 8.75% occurred between populations. The genetic differentiation value of 0.083 was within the range of moderate differentiation. The STRUCTURE analysis showed that ΔK reached its maximum value (ΔK = 310.98) at K = 2 ( Supplementary Figure S3), suggesting that the 150 individuals from the five populations most likely belong to two principal genetic clusters (Figure 2(A)). Cluster one mainly included individuals from the WBL, WJL, and XCW populations, whereas cluster two mainly included individuals from the XL and ZWL populations. The STRUCTURE results at K = 3 were consistent with those of the NJ tree (Figure 2(B)), and there were mixed individuals among the populations, indicating some gene flow or differentiation.

Table 2.

Analysis of Molecular Variance (AMOVA) based on 17 EST-SSR Markers for Populations of R. longipedicellatum.

Figure 2.

Results of Bayesian clustering of 150 individuals of R. longipedicellatum. Note: (A) Assignment of 150 individuals into K = 2 genetically distinguishable groups. (B) Assignment of 150 individuals into K = 3 genetically distinguishable groups. Each individual is represented by a vertical bar colored according to the assigned group(s).

10.1177_19400829221078112-fig2.tif

Based on the DAS genetic distance, NJ trees of populations and individuals were constructed, as shown in Figure 3. The genetic relationship between the WJL and XCW populations was close, clustering as a clade, and the XL and ZWL populations clustered into another clade. The WBL population formed a separate clade and belonged to a large branch along with the WJL and XCW populations (Figure 3(A)). The NJ tree of the 150 individuals (Figure 3(B)) revealed some cross-mixing among the three groups and among the populations. The PCoA analysis of the five populations is shown in Figure 4. Approximately 66.01% of the variation occurred between principal coordinates one and two, and their contribution rates were 42.47% and 23.54%, respectively. The results were similar to the NJ tree, and the five populations were roughly divided into two groups. The genetic distance and geographical distance among the five populations were positively correlated (Mantel test, r = 0.77, t = 2.95, p < 0.01). This result is consistent with the NJ tree and two-dimensional PCoA map.

Figure 3.

Neighbor-Joining (NJ) dendrogram of five extant populations and 150 individuals of R. longipedicellatum based on DAS genetic distance. The bootstrap values are indicated at the nodes in each cluster.

10.1177_19400829221078112-fig3.tif

Figure 4.

Principal coordinates analysis (PCoA) of five R. longipedicellatum populations based on a genetic distance matrix of 17 EST-SSR loci. Note: The first and second axes extracted 42.47% and 23.54% of the total genetic variance, respectively.

10.1177_19400829221078112-fig4.tif

Bottleneck Effect Analysis

The results of the bottleneck analysis for the five residual populations based on the SMM and TPM are shown in Table 3. Assuming that all of the EST-SSR loci conformed to the SMM model, the Sign test and Wilcoxon test detection methods showed that all five populations experienced heterozygote deficit or excess and a certain degree of genetic bottleneck. Based on the TPM model, five populations all experienced heterozygote excess, and the ZWL population showed a significant deviation from mutation-drift equilibrium in Wilcoxon test (p < 0.01, after Bonferroni correction). The G-W values of the five populations ranged from 0.22 to 0.26 with an overall value of 0.26 (Table 3), which was lower than the critical value of 0.68 (Garza & Williamson, 2001) and consistent with a population decline in all five populations. At the species level, the results of the different detection methods consistently indicated a genetic bottleneck.

Table 3.

Bottleneck Analysis for Five Populations of R. longipedicellatum.

Discussion

Genetic Diversity

The evolutionary potential of a species and its ability to withstand adverse environments depend on both the genetic diversity of the species and its population genetic structure (Stebbins, 1950). Different levels of genetic variation and genetic structure of populations can be attributed to multiple factors, such as breeding systems, biological characteristics, evolutionary history, and natural selection (Hamrick & Godt, 1996; Oyundelger et al., 2021; Zhou et al., 2020). The results of this EST-SSR study showed that R. longipedicellatum has high genetic diversity at the species level (He = 0.559, NA = 9.529), and the mean genetic diversity of the five populations investigated (He = 0.507, NA = 5.910) was higher than that of Rhododendron oldhamii (He = 0.284, ranged 0.240–0.336) (Hsieh et al., 2013) and the Rhododendron pseudochrysanthum complex (average He = 0.424) (Chen et al., 2014), but lower than the genetic diversity of Rhododendron jinggangshanicum (He = 0.642) (M. Li et al., 2015) and Rhododendron decorum (He = 0.758) (Wang et al., 2013) based on gSSR markers. Although gSSR markers vary more, possibly due to a higher mutation rate, the variation observed in EST-SSR markers may be more functionally significant and represent responses to selection more effectively (Ellis & Burke, 2007; Kane & Rieseberg, 2007).

Despite the restricted distribution and small population sizes of R. longipedicellatum, this species has retained high levels of genetic diversity, possibly due to the following reasons. The extensive temporal and spatial heterogeneity of the limestone habitat in which these plants occur may contribute to their genetic diversity. The limestone habitat includes variation in light, temperature, rock type, degree of differentiation, fissure soil, karst water, humus layer thickness, and the species composition and density of the vegetation. Some R. longipedicellatum plants occurring in the limestone habitat are attached to tree trunks, providing a stable yet polymorphic microenvironment that reduces the impact of natural and human-induced disturbance. Second, outcrossing plants tend to have higher genetic variation than selfing plants (Huang et al., 2021; Nybom, 2004). The cross-pollination reproductive system of R. longipedicellatum reported by Li, Liu, Li, Ma, et al. (2018) may, therefore, also support the relatively high genetic diversity in this species. The high genetic variability of R. longipedicellatum may also be attributed to abundant diversity in the ancestral population (Zhang et al., 2021) that may have existed as a large, continuous, and widespread species, although little is known about the origin and evolutionary history of R. longipedicellatum. The current restricted distribution and small population sizes of the species result from habitat deterioration and loss, as well as direct human interference. R. longipedicellatum is a perennial woody plant whose biological characteristics may also play a role in the maintenance of genetic diversity. In the five populations of R. longipedicellatum studied, few or no seedlings were observed and the plants usually propagated vegetatively, which tends to maintain genetic diversity.

Genetic Structure

According to AMOVA, 91.25% of the total genetic variation occurred within populations, and only 8.75% was inter-population variation, which is consistent with the observation that most of the genetic variation in woody plants occurs within populations (Liu et al., 2012; X. Liu et al., 2020). In this study, the overall FST was 0.083, indicating moderate differentiation among populations (Jakobsson et al., 2013) due to drift or possibly divergent selection among patchily-distributed populations caused by human disturbance or habitat heterogeneity (Liu et al., 2019; Willi et al., 2006). The STRUCTURE analysis divided 150 individuals from the five existing populations into two genetic populations, and the NJ tree and PCoA analyses also divided the populations into two groups. This clustering of populations may be due to local environmental conditions, which often shape patterns of genetic structure, especially in heterogeneous and fragmented habitats (Young et al., 1996; Zhu et al., 2016). Both at the species and the population levels, FIS was less than 0 and the cross-pollination rate was greater than 1, indicating that outbreeding was present in the population, contrary to the high inbreeding coefficient expected in a small population of plants occurring in a unique habitat and subjected to an increasingly shrinking geographical distribution. The obvious homozygote deficiency or heterozygote excess may be associated with population regeneration in some plants (Stoeckel et al., 2010). In addition, the increased heterozygosity may be due to the small number of populations with few individual plants in each population or to the presence of a high allelic richness in the parent plants (Huang et al., 2021; Pudovkin et al., 1996).

The magnitude of gene flow between populations has a significant influence on the genetic structure of populations and is a key factor that predicts the effect of changes in the environment, such as human interference and population isolation (Wu et al., 2017). In this study, the estimated number of migrants per generation among populations of R. longipedicellatum was >1, indicating that gene flow was high enough to overcome differentiation among populations by genetic drift (Wright, 1949). The diffusion and transmission of pollen and seeds are the two major forms of gene exchange between populations. Similar to the majority of Rhododendron species (Ng & Corlett, 2000; Yang et al., 2020), our field surveys found that the seed transmission capability of R. longipedicellatum was very limited, seedlings in the populations were rare, and the pollen dispersal depended mainly on male Bombus braccatus (Li, Liu, Li, Ma, et al., 2018), with a propagation distance of 3–10 km (Ng & Corlett, 2000). All of the study populations of R. longipedicellatum were geographically close, and even the relatively distant populations, WBL and ZWL, were only about 9 km apart. Thus, pollen flow by pollinators is expected.

Population Bottlenecks

Bottleneck v1.2 identified genetic bottlenecks in R. longipedicellatum using different models and detection methods. The results showed that at the species level, under the assumptions of the SMM and TPM models, R. longipedicellatum significantly deviated from mutation-drift equilibrium, and at the population level, the ZWL population also deviated from equilibrium (p < 0.01, after Bonferroni correction). Recent damage caused by human activities, such as grazing and indiscriminate felling, has increased, especially in the ZWL population. The M-ratio test reflected a loss of allelic diversity in all five populations, consistent with population contraction and supporting the Bottleneck results. As bottlenecks tend to cause reduction the quality of the species gene pool and genetic variability of populations and consequently, their fitness. The high genetic diversity observed in R. longipedicellatum in this study may be due to the occurrence of the genetic bottleneck being relatively recent.

Implications for Conservation

These results indicate that genetic diversity is still present, to a certain degree, in R. longipedicellatum. To preserve this species, an effective conservation measure could be to strengthen the habitat protection of R. longipedicellatum, strictly forbid indiscriminate felling, and protect it from man-made interference. All five populations contained a high level of genetic diversity, with the XCW population having the richest genetic diversity and the highest number of effective alleles. The connectivity and mixing between XCW and other populations were higher than those between XCW and other populations, which indicated that XCW is of great significance in maintaining genetic polymorphism. This population is geographically located in the middle of the other populations and thus beneficial for facilitating dispersal and connecting the other populations as a corridor. Therefore, we suggest that additional conservation efforts be focused on the XCW population. In addition, protection efforts can be enhanced by appealing to the related local authorities, setting up conservation warning signs, and publicizing the value of, and conservation strategy for, this endangered species to local residents.

Acknowledgments

We would like to thank all the funding for their support. We would also like to thank Editage ( www.editage.com) for English language editing. Finally, we are particularly grateful to the anonymous referees for their time and their valuable comments.

Declaration of Conflicting Interests The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

FundingThe author(s) disclosed receipt of the following financial support for the research, authorship, and/or publication of this article: This work was supported by the Fundamental Research Funds for the Central Non-profit Research Institution of the Chinese Academy of Forestry (CAFYBB2019ZB007); the Biodiversity Survey and Assessment Project of the Ministry of Ecology and Environment, China (2019HJ2096001006); the Ten Thousand Talent Program of Yunnan Province (YNWR-QNBJ-2019-010; YNWR-QNBJ-2018-174) and the Young and middle-aged academic and technical leaders in Yunnan Province reserve talents (2018HB066).

Supplemental materialSupplemental material for this article is available online.

References

1.

Cai , L. , Neilsen , J. , Dao , Z. L. , Ma , Y. P. (2016). Rhododendron longipedicellatum (Ericaceae), a new species from Southeastern Yunnan, China. Phytotaxa, 282(4), 296–300. DOI: https://doi.org/10.11646/phytotaxa.282.4.7.  https://doi.org/10.11646/phytotaxa.282.4.7Google Scholar

2.

Chen , C. Y. , Liang , B. K. , Chung , J. D. , Chang , C. T. , Hsieh , Y. C. , Lin , T. C. , Hwang , S. Y. (2014). Demography of the upward-shifting temperate woody species of the Rhododendron pseudochrysanthum complex and ecologically relevant adaptive divergence in its trailing edge populations. Tijdschrift voor Gerontologie en Geriatrie, 10(1), 111–126.  https://doi.org/10.1007/s11295-013-0669-xGoogle Scholar

3.

Chybicki , I. J. , Burczyk , J. (2009). Simultaneous estimation of null alleles and inbreeding coefficients. Journal of Hered, 100(1), 106–113.  https://doi.org/10.1093/jhered/esn088Google Scholar

4.

Cornuet , J. M. , Luikart , G. (1997). Description and power analysis of two tests for detecting recent population bottlenecks from allele frequency data. Genetics, 144(4), 2001–2014.  https://doi.org/10.1093/genetics/144.4.2001Google Scholar

5.

Dakin , E. E. , Avise , J. C. (2004). Microsatellite null alleles in parentage analysis. Heredity, 93(5), 504–509.  https://doi.org/10.1038/sj.hdy.6800545Google Scholar

6.

Doyle , J. , Doyle , J. (1987). A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem Bull, 19(1), 11–15.  https://doi.org/10.12691/ajmr-3-1-7Google Scholar

7.

Earl , D. A. , VonHoldt , B. M. (2012). STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genetics Resources, 4(2), 359–361.  https://doi.org/10.1007/s12686-011-9548-7Google Scholar

8.

Ellis , J. R. , Burke , J. M. (2007). EST-SSRs as a resource for population genetic analyses. Heredity, 99(2), 125–132.  https://doi.org/10.1038/sj.hdy.6801001Google Scholar

9.

Evanno , G. , Regnaut , S. , Goudet , J. (2005). Detecting the number of clusters of individuals using the software structure: a simulation study. Molecular Ecology, 14(8), 2611–2620.  https://doi.org/10.1111/j.1365-294X.2005.02553.xGoogle Scholar

10.

Excoffier , L. , Lischer , H. E. (2010). Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Molecular Ecology Resources, 10(3), 564–567.  https://doi.org/10.1111/j.1755-0998.2010.02847.xGoogle Scholar

11.

Frankham , R. (2003). Genetics and conservation biology. Comptes Rendus Biologies, 326(1), 22–29.  https://doi.org/10.1016/S1631-0691(03)00023-4Google Scholar

12.

Garza , J. C. , Williamson , E. G. (2001). Detection of reduction in population size using data from microsatellite loci. Molecular Ecology, 10(2), 305–318.  https://doi.org/10.1046/j.1365-294X.2001.01190.xGoogle Scholar

13.

Goudet , J. (2001). FSTAT, a program to estimate and test gene diversities and fixation indices.  http://www.unil.ch/izea/softwares/FSTat.html. Version 2.9.3 Google Scholar

14.

Hamrick , J. L. , Godt , M. J. W. (1996). Effects of life history traits on genetic diversity in plant species. Philosophical Transactions of the Royal Society B-Biological Sciences, 351, 1291–1298.  http://www.jstor.org/stable/56204Google Scholar

15.

Hsieh , Y. C. , Chung , J. D. , Wang , C. N. , Chang , C. T. , Chen , C. Y. , Hwang , S. Y. (2013). Historical connectivity, contemporary isolation and local adaptation in a widespread but discontinuously distributed species endemic to Taiwan, Rhododendron oldhamii (Ericaceae). Heredity, 111(2), 147–156.  https://doi.org/10.1038/hdy.2013.31Google Scholar

16.

Huang , R. , Wang , Y. , Li , K. , Wang , Y. Q. (2021). Genetic variation and population structure of clonal zingiber zerumbet at a fine geographic scale: a comparison with two closely related selfing and outcrossing zingiber species. BMC Ecology Evolution, 21(1), 116.  https://doi.org/10.1186/s12862-021-01853-2Google Scholar

17.

IUCN (2001) IUCN red list Categories and Criteria: version 3.1 (30). IUCN Species Survival Commission. Google Scholar

18.

Jakobsson , M. , Edge , M. D. , Rosenberg , N. A. (2013). The relationship between FST and the frequency of the most frequent allele. Genetics, 193(2), 515–528.  https://doi.org/10.1534/genetics.112.144758Google Scholar

19.

Kane , N. C. , Rieseberg , L. H. (2007). Selective sweeps reveal candidate genes for adaptation to drought and salt tolerance in common sunflower, Helianthus annuus. Genetics, 175(4), 1823–1834.  https://doi.org/10.1534/genetics.106.067728Google Scholar

20.

Langella , O. (2007). Population genetic software (individuals or populations distances, phylogenetic trees). Populations 1.2.30. Available online at: http://bioinformatics.org/∼tryphon/populations/Google Scholar

21.

Li , M. W. , Chen , S. F. , Shi , S. , Zhang , Z. , Liao , W. B. , Wu , W. , Zhou , R. C. , Fan , Q. (2015). High genetic diversity and weak population structure of Rhododendron jinggangshanicum, a threatened endemic species in Mount Jinggangshan of China. Biochemical Systematics & Ecology, 58, 178–186.  https://doi.org/10.1016/j.bse.2014.12.008Google Scholar

22.

Li , T. Q. , Liu , X. F. , Li , Z. H. , Ma , H. , Wan , Y. M. , Liu , X. X. , Fu , L. Y. (2018). Study on reproductive biology of Rhododendron longipedicellatum: A newly discovered and special threatened plant surviving in limestone habitat in Southeast Yunnan, China. Facial Plastic Surgery: FPS, 9, 33.  https://doi.org/10.3389/fpls.2018.00033Google Scholar

23.

Li , T. Q. , Liu , X. F. , Li , Z. H. , Wan , Y. M. , Liu , X. X. , Ma , H. (2018). Development of novel EST-SSR markers for Rhododendron longipedicellatum (Ericaceae) and cross-amplification in two congeners. Applications in Plant Sciences, 6(6), e01162.  https://doi.org/10.1002/aps3.1162Google Scholar

24.

Li , Y. , Zhang , X. W. , Fang , Y. M. (2019). Landscape features and climatic forces shape the genetic structure and evolutionary history of an oak species (quercus chenii) in east china. Frontiers in Plant Science, 10, 1060.  https://doi.org/10.3389/fpls.2019.01060Google Scholar

25.

Liu , X. F. , Ma , Y. P. , Wan , Y. M. , Li , Z. H. , Ma , H. (2020). Genetic diversity of Phyllanthus emblica from two different climate type areas. Facial Plastic Surgery: FPS, 11, 580812.  https://doi.org/10.3389/fpls.2020.580812Google Scholar

26.

Liu , Y. L. , Geng , Y. P. , Song , M. L. , Zhang , P. F. , Hou , J. L. , Wang , W. Q. (2019). Genetic structure and diversity of Glycyrrhiza populations based on transcriptome SSR markers. Plant Molecular Biology Reporter, 37(5-6), 401–412.  https://doi.org/10.1007/s11105-019-01165-2Google Scholar

27.

Liu , Y. F. , Xing , M. , Zhao , W. , Fan , R. J. , Luo , S. , Chen , X. (2012). Genetic diversity analysis of Rhododendron aureum Georgi (Ericaceae) located on Changbai Mountain using ISSR and RAPD markers. Plant Systematics & Evolution, 298(5), 921–930.  https://doi.org/10.1007/s00606-012-0601-0Google Scholar

28.

Ma , H. , Li , T. Q. , Liu , X. F. , Wan , Y. M. , Li , Y. Y. , Liu , X. X. , Li , Z. H. (2017). Research progress in conservation biology of Rhododendron. World Forestry Research, 30(4), 13–17.  https://doi.org/10.13348/j.cnki.sjlyyj.2017.0037.yGoogle Scholar

29.

Ma , H. , Liu , Y. B. , Liu , D. T. , Sun , W. B. , Liu , X. F. , Wan , Y. M. , Zhang , X. J. , Zhang , R. G. , Yun , Q. Z. , Wang , J. H. , Li , Z. H. , Ma , Y. P. (2021). Chromosome-level genome assembly and population genetic analysis of a critically endangered rhododendron provide insights into its conservation. The Plant Journal, 107(5), 1533–1545.  https://doi.org/10.1111/tpj.15399Google Scholar

30.

Ma , Y. P. , Nielsen , J. , Chamberlain , D. F. , Li , X. Y. , Sun , W. B. (2014). The conservation of Rhododendrons is of greater urgency than has been previously acknowledged in China. Biodiversity & Conservation, 23(12), 3149–3154.  https://doi.org/10.1007/s10531-014-0764-9Google Scholar

31.

Mable , B. K. (2019). Conservation of adaptive potential and functional diversity: integrating old and new approaches. Conservation Genetics, 20(1), 89–100.  https://doi.org/10.1007/s10592-018-1129-9Google Scholar

32.

Nei , M. (1973). Analysis of gene diversity in subdivided populations. Proceedings of the National Academy of Sciences of the United States of America, 70(12), 3321–3323.  https://doi.org/10.1073/pnas.70.12.3321Google Scholar

33.

Ng , S. C. , Corlett , R. T. (2000). Comparative reproductive biology of the six species of Rhododendron (Ericaceae) in Hong Kong, South China. Industrial Crops & Products, 78(2), 221–229.  https://doi.org/ 10.1139/b99-181Google Scholar

34.

Nybom , H. (2004). Comparison of different nuclear DNA markers for estimating intraspecific genetic diversity in plants. Molecular Ecology, 13(5), 1143–1155.  https://doi.org/10.1111/j.1365-294X.2004.02141.xGoogle Scholar

35.

Oyundelger , K. , Herklotz , V. , Harpke , D. , Oyuntsetseg , B. , Ritz , C. M. (2021). Contrasting effects of local environment and grazing pressure on the genetic diversity and structure of Artemisia frigida. Conservation Genetics, 39(2), 364–375.  https://doi.org/10.1007/s10592-021-01375-wGoogle Scholar

36.

Paradis , E. , Schliep , K. (2019). ape 5.0: An environment for modern phylogenetics and evolutionary analyses in R. Bioinformatics, 35(3), 526–528.  https://doi.org/10.1093/bioinformatics/bty633Google Scholar

37.

Piry , S. , Luikart , G. , Cornuet , J.-M. (1999). Bottleneck: A computer program for detecting recent reductions in the effective population size using allele frequency data. Journal of Heredity, 90(4), 502–503.  https://doi.org/10.1093/jhered/90.4.502Google Scholar

38.

Pritchard , J. K. , Stephens , M. , Donnelly , P. (2000). Inference of population structure using multilocus genotype data. Genetics, 155(2), 945–959.  https://pubmed.ncbi.nlm.nih.gov/10835412/Google Scholar

39.

Pudovkin , A. I. , Zaykin , D. V. , Hedgecock , D. (1996). On the potential for estimating the effective number of breeders from heterozygote-excess in progeny. Genetics, 144(1), 383–387.  https://doi.org/10.1093/genetics/144.1.383Google Scholar

40.

Rice , W. R. (1989). Analyzing tables of statistical tests. Evolution, 43(1), 223–225.  https://doi.org/10.1111/j.1558-5646.1989.tb04220.xGoogle Scholar

41.

Singh , R. K. , Jena , S. N. , Khan , S. , Yadav , S. , Banarjee , N. , Raghuvanshi , S. , Bhardwaj , V. , Dattamajumder , S. K. , Kapur , R. , Solomon , S. , Swapna , M. , Srivastava , S. , Tyagi , A. K. , Tyagi , A. K. (2013). Development, cross-species/genera transferability of novel EST-SSR markers and their utility in revealing population structure and genetic diversity in sugarcane. Genes, 524(2), 309–329.  https://doi.org/10.1016/j.gene.2013.03.125Google Scholar

42.

Slatkin , M. (1985). Gene flow in natural populations. Annual Review of Ecology Evolution & Systematics, 16, 393–430.  https://doi.org/10.1146/annurev.es.16.110185.002141Google Scholar

43.

Stebbins , C. L. J (1950). Variation & evolution in plants. Oxford University Press. Google Scholar

44.

Stoeckel , S. , Grange , J. , Fernández-Manjarres , J. F. , Bilger , I. , Frascaria-Lacoste , N. , Mariette , S. (2010). Heterozygote excess in a self-incompatible and partially clonal forest tree species - Prunus avium L. Molecular Ecology, 15(8), 2109–2118.  https://doi.org/10.1111/j.1365-294X.2006.02926.xGoogle Scholar

45.

Wang , X. Q. , Huang , Y. , Long , C. L. (2013). Assessing the genetic consequences of flower-harvesting in Rhododendron decorum Franchet (Ericaceae) using microsatellite markers. Biochemical Systematics & Ecology, 50, 296–303.  https://doi.org/10.1016/j.bse.2013.04.009Google Scholar

46.

Wei , W. L. , Qi , X. Q. , Wang , L. H. , Zhang , Y. X. , Hua , W. , Li , D. H. , Lv , H. X. , Zhang , X. R. (2011). Characterization of the sesame (Sesamum indicum L.) global transcriptome using illumina paired-end sequencing and development of est-ssr markers. BMC Genomics, 12(1), 451.  https://doi.org/10.1186/1471-2164-12-451Google Scholar

47.

Weir , B. S. , Cockerham , C. C. (1984). Estimating F-statistics for the analysis of population structure. Evolution, 38, 1358–1370.  https://doi.org/10.1111/j.1558-5646.1984.tb05657.xGoogle Scholar

48.

Willi , Y. , Buskirk , J. V. , Hoffmann , A. A. (2006). Limits to the adaptive potential of small populations. Annual Review of Ecology Evolution & Systematics, 37(1), 433–458.  https://doi.org/10.2307/30033839Google Scholar

49.

Wright , S. (1949). The genetic structure of populations. Annals of Human Genetics, 15(1), 323–354.  https://doi.org/10.1111/j.1469-1809.1949.tb02451.xGoogle Scholar

50.

Wu , F. Q. , Shen , S. K. , Zhang , X. , Yang , G. , Wang , Y. H. (2017). Inferences of genetic structure and demographic history of Rhododendron protistum var. giganteum-The world’s largest Rhododendron using microsatellite markers. Flora, 233, 1–6.  https://doi.org/10.1016/j.flora.2017.04.009Google Scholar

51.

Yadav , L. K. , Mcassey , E. V. , Wilde , H. D. (2019). Genetic diversity and population structure analysis of Rhododendron canescens, a Native Azalea for urban landscaping. HortScience, 54(4), 647–651.  https://doi.org/10.21273/HORTSCI13840-18Google Scholar

52.

Yang , Y. , Wang , H. Y. , Ma , L. H. (2020). Characteristics of fruiting and seed germination of endangered plant, Rhododendron changii. Chinese Journal of Scientia Silvae Sinicae, 56(10), 173–183.  https://doi.org/10.11707/j.1001-7488.20201019Google Scholar

53.

Yeh , F. C. , Yang , R. C. , Boyle , T. (1999). POPGENE VERSION 1.31: Microsoft Window–based Freeware for Population Genetic Analysis.  ftp://ftp.microsoft.com/Softlib/HPGLGoogle Scholar

54.

Young , A. , Boyle , T. , Brown , T. (1996). The population genetic consequences of habitat fragmentation for plants. Trends in Ecology & Evolution, 11, 413–418.  https://doi.org/10.1016/0169-5347(96)10045-8Google Scholar

55.

Zhang , X. J. , Liu , X. F. , Liu , D. T. , Cao , Y. R. , Li , Z. H. , Ma , Y. P. , Ma , H. (2021). Genetic diversity and structure of Rhododendron meddianum, a plant species with extremely small populations. Plant Diversity, 43(6), 472–479.  https://doi.org/10.1016/j.pld.2021.05.005Google Scholar

56.

Zhang , Y. , Zhang , X. , Wang , Y. H. , Shen , S. K. (2017). De Novo Assembly of transcriptome and development of novel EST-SSR markers in Rhododendron rex Lévl. through Illumina sequencing. Facial Plastic Surgery: FPS, 8, 1664.  https://doi.org/10.3389/fpls.2017.01664Google Scholar

57.

Zhou , Q. , Mu , K. , Ni , Z. X. , Liu , X. H. , Li , Y. A. , Xu , L. A. (2020). Analysis of genetic diversity of ancient Ginkgo populations using SSR markers. Industrial Crops & Products, 145, e111942.  https://doi.org/10.1016/j.indcrop.2019.111942Google Scholar

58.

Zhu , Z. , Zhang , L. , Gao , L. , Tang , S. , Zhao , Y. , Yang , J. (2016). Local habitat condition rather than geographic distance determines the genetic structure of Tamarix chinensis populations in Yellow River Delta, China. Tijdschrift Voor Gerontologie en Geriatrie, 12(1), 14.  https://doi.org/10.1007/s11295-016-0971-5Google Scholar
© The Author(s) 2022
Yurong Cao, Yongpeng Ma, Zhenghong Li, Xiongfang Liu, Detuan Liu, Suping Qu, and Hong Ma "Genetic Diversity and Population Structure of Rhododendron longipedicellatum, an Endangered Species," Tropical Conservation Science 15(1), (11 March 2022). https://doi.org/10.1177/19400829221078112
Received: 21 April 2021; Accepted: 18 January 2022; Published: 11 March 2022
KEYWORDS
bottleneck effect
conservation genetics
endangered species
EST-SSR
molecular markers
Back to Top