经济文库 - 千万精品文档,你想要的都能搜到,下载即用。

生命科学学院-李晶.pdf

瑾年 Invader17 页 1.953 MB 访问 2872.97下载文档
生命科学学院-李晶.pdf生命科学学院-李晶.pdf生命科学学院-李晶.pdf生命科学学院-李晶.pdf生命科学学院-李晶.pdf生命科学学院-李晶.pdf
当前文档共17页 2.97
下载后继续阅读

生命科学学院-李晶.pdf

Received: 9 October 2018 | Revised: 12 April 2019 | Accepted: 15 April 2019 DOI: 10.1002/ece3.5278 ORIGINAL RESEARCH Phylogeography of Dendrolimus punctatus (Lepidoptera: Lasiocampidae): Population differentiation and last glacial maximum survival Jing Li1 | Qian Jin1,2 | Geng‐ping Zhu3 1 | Chong Jiang1 | Ai‐bing Zhang1 College of Life Sciences, Capital Normal University, Beijing, China Abstract 2 Although the Masson pine moth, Dendrolimus punctatus, is one of the most de‐ Suqian Institute of Agricultural Sciences, Jiangsu Academy of Agricultural Sciences, Suqian, China 3 College of Life Sciences, Tianjin Normal University, Tianjin, China Correspondence Ai‐bing Zhang, College of Life Sciences, Capital Normal University, Beijing, China. Email: zhangab2008@cnu.edu.cn Funding information China National Funds for Distinguished Young Scientists, Grant/Award Number: 31425023; Natural Science Foundation of China, Grant/Award Number: 31601877, 31772501 and 31400191; Beijing Municipal Natural Science Foundation, Grant/Award Number: 5172005 structive forest pest insects and is an endemic condition in China, we still do not fully understand the patterns of how its distribution range varies in response to Quaternary climatic oscillations. Here, we sequenced one maternally inherited mitochondrial gene (COI) and biparentally inherited nuclear data (ITS1 and ITS2) among 23 natural populations across the entire range of the species in China. A total of 51 mitotypes and 38 ribotypes were separately obtained using mtDNA and ITS1 data. Furthermore, significant phylogeographical structure (NST > GST, p < 0.01) were detected. The spatial distribution of mitotypes implied that two distinct groups existed in the species: one in the southwest distribution, including 10 locations, and the other located in the northeast region of China. It is suggested, therefore, that each group was derived from ancestors that occupied different iso‐ lated refugia during previous periods, possibly last glacial maximum. Mismatch dis‐ tribution and Bayesian population dynamics analysis suggested the population size underwent sudden expansion, which is consistent with the results of ecological niche modeling. As a typical phytophagous insect, the history of population expan‐ sion was in accordance with the host plants, providing abundant food resources and habitat. Intraspecific success rate of barcoding identification was lower than interspecific ones, indicating a level of difficulty in barcoding individuals from dif‐ ferent populations. However, it still provides an early insight into the pattern of genetic diversity within a species. KEYWORDS Dendrolimus punctatus, DNA barcode, phylogeography, population expansion, quaternary climate oscillation Jing Li and Qian Jin contributed equally to this work. This is an open access article under the terms of the Creat​ive Commo​ns Attri​bution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited. © 2019 The Authors. Ecology and Evolution published by John Wiley & Sons Ltd. 7480 | www.ecolevol.org  Ecology and Evolution. 2019;9:7480–7496. | 7481 LI et al. 1 | I NTRO D U C TI O N It has been widely accepted that the distribution and genetic struc‐ ture of current organism have been greatly affected by environ‐ mental factors and their historical processes (Hewitt, 2000). The oscillations of Quaternary climate changes played critical roles in multiple contraction–expansion processes, which have greatly shaped current geographical distributions and genetic differentia‐ tions of many species in temperate zones of East Asian (Hickerson et al., 2010; Qiu, Fu, & Comes, 2011). However, compared with the numerous studies focused in the Qinghai‐Tibet Plateau and Mts. Hengduan (Lei, Qu, Song, Alström, & Fjeldså, 2015; Liu et al., 2016; Wang et al., 2018), such investigations have just started in the sub‐ tropical regions of China. During the Last Glacial Maximum (LGM) period, the climate of this region became much cooler and dryer, F I G U R E 1 Photograph of Dendrolimus punctatus (Walker, 1985) (by Dr. Chong‐Hui Yang) which drove the southward migration of the warm‐temperate ev‐ ergreen forests as far as c.24°N (Harrison, Yu, Takahara, & Prentice, COI gene as well as two nuclear loci: ribosomal first internal 2001; Qiu et al., 2011). Moreover, the vast continental hilly areas in transcribed spacer (rDNA ITS1) and second spacer (rDNA ITS2). the subtropical China provided potential refugia for temperate de‐ Two different inherited molecular markers were included in this ciduous forests and boreal conifer trees (Jiang et al., 2016; Lei et study: mtDNA marker (COI) represented maternal inheritance; al., 2015). the ITS fragments represented the biparental inheritance. In Previous research implied that much of the genetic variety of addition, these three molecular markers shared barcoding se‐ phytophagous insects was motivated by the diverging specializa‐ quences well and proved to be excellent methods to help ex‐ tion processes onto different host plants (Peterson & Denno, 1998). plain the interspecific relationship among the sibling species of More detailed studies based on population‐level variations of host D. punctatus (Dai et al., 2012). Nevertheless, although DNA bar‐ plants implied that the diversification of host could have an influence coding has been widely used as a concept to facilitate biological on the differentiation of insects (Friberg, Schwind, & Roark, 2014; identifications at the species level (Hebert, Cywinska, Ball, & Friberg, Schwind, & Thompson, 2016; Matsubayashi, Ohshima, deWaard, 2003), there has been limited use related to lower cat‐ & Nosil, 2010; Singer, Ng, & Moore, 1991). Dendrolimus punctatus egories such as subspecies and populations (Huemer & Hebert, (Walker, 1985) distributed in subtropical China was adopted in our 2011; Valade et al., 2009). These lower categories, defined by study to identify whether the genetic differentiation of host plants, evidence of geographical, morphological, and ecological criteria, caused by climate changes during the ice age, affects genetic vari‐ can be potential or cryptic species (DeSalle, 2006). It would be ation of their predators, the phytophagous insects. Dendrolimus particularly desirable to test whether these widely accepted in‐ species (Lepidoptera: Lasiocampidae) are one of the most serious terspecific barcodes could also be adopted as population identi‐ phytophagous pests worldwide and caused extensive forest damage fication. Although the typical DNA barcoding was not sufficient (Dai et al., 2012; Zhao, Wu, Lv, Chen, & Lin, 1999; Zhang, Wang, Tan, to identify geographical clusters within species (Hajibabaei, & Li, 2003; Zhang, Li, Chen, & Zhang, 2004) (Figure 1). Dendrolimus Singer, Hebert, & Hickey, 2007), the limited geographical iso‐ punctatus is endemic to southern Asia and, in recent years, has be‐ lation was observed between different regions (Rach, DeSalle, come the most serious and economically damaging insect pest in the Sarkar, Schierwater, & Hadrys, 2008; Zhao et al., 1999). The suc‐ south China forests. Pinus massoniana, which is a typical warm‐tem‐ cess rate of barcoding at lower categories largely depends on perate evergreen coniferous in south mainland China, is the main the phylogeographic structure of the taxon (Zhang et al., 2003). host plant of D. punctatus (Cai, 1995; Hou, 1987). Based on a phylo‐ Factors such as the geographic distribution of the populations, geographic analysis, P. massoniana experienced a recent expansion time and degree of isolation, effective population sizes, and par‐ of its smaller populations during the Quaternary period (Ge et al., ticularly gene flow showed fundamental effects. Therefore, our 2012). The increasing population density, which increased rapidly present work seeks to address the following points: after a few generations, posed a great threat to forestry production. Consequently, great efforts were devoted to keep the size of popu‐ lation at a relatively low and stable level. Notably, low‐density pop‐ ulations are difficult to detect in the field, which makes it extremely arduous to collect samples. Here, we examined the genetic variation across the distri‐ bution areas of D. punctatus with the barcode fragment of the 1. to investigate the phylogeographical structure of the masson pine pest; 2. to test whether the demographic history of this moth corresponds to the “contraction‐expansion” mode similar to its host plant; 3. to explore the potential availability of population identification using the DNA barcode. 7482 | LI et al. F I G U R E 2 A map of the sampling sites and the geographic distribution of Dendrolimus punctatus mtDNA haplotypes. Pie charts show the proportion of mitotypes within each population, and colors represent different lineages. The numbers beside the circles represent population numbers listed in Table 1 2 | M ATE R I A L S A N D M E TH O DS 2.1 | Specimen sampling ITS2), and the COI barcodes were also submitted to the BOLD system (accession numbers: GBGL14669–GBGL14904). gathered with tweezers during the day. The larvae of D. punctatus dis‐ 2.3 | Phylogenetic analysis and network reconstruction played an aggregated distribution pattern. Hence, one larva of pine The raw DNA sequences were all checked manually and aligned The specimens of D. punctatus were collected as larvae, which were caterpillar was collected every 5 m. The sampling scheme, including a total of 236 specimens from 23 localities, covered the major geo‐ graphic distribution of this species (Figure 2; Table 1). The specimens were collected in the natural forests. In order to preserve DNA se‐ quence integrity, the larvae collected in natural forests were imme‐ diately placed in 95% ethanol and ultimately stored at 4°C. Voucher specimens were deposited at the Capital Normal University, Beijing, China. with software MUSCLE (Edgar, 2004). Aligned COI sequences were translated into amino acid sequences using the MEGA program (Kumar, Tamura, & Nei, 2004). The phylogenetic relationships within D. punctatus were reconstructed using maximum likelihood (ML) in‐ ference and Bayesian inference (BI). Two specimens from species Dendrolimus superans (Butler, 1877) were used as outgroups. The best‐fit model of nucleotide substitution was selected from 88 models using the Akaike information criterion (AIC) with jModel‐ Test 2 (Darriba, Taboada, Doallo, & Posada, 2012). The ML analysis 2.2 | DNA extraction, PCR amplification, and sequencing was conducted using the program RAxML v.7.0.4 with 1,000 boot‐ DNA samples were isolated from the legs of adult individuals using v3.1.2 (Ronquist & Huelsenbeck, 2003). The MCMC analysis was DNeasy Blood and Tissue kit (Biomed) following the manufacturer's protocol. Sequences of mitochondrial gene COI and nuclear ITS were amplified using rTaq (Takara). The same PCR primers and amplification reaction of Dai et al. (2012) were used in our study. The amplification products were subjected to electrophoresis in a 2% (w/v) agarose gel in TAE buffer (0.04 M Tris–acetate, 0.001 M EDTA) to check whether the amplification reactions were successful. Then, the amplification prod‐ ucts were used for sequencing, which was performed with an ABI3130 sequencer at Biomed Biotechnology Co., Ltd. All sequences have been straps conducted (Alexandros, 2006). The GTR+G model was used for all genes. The Bayesian analysis was performed using MrBayes run for 10,000,000 generations, following a burn‐in series of 1,000 generations. Haplotype networks for COI sequence and each ITS dataset were constructed to better visualize the reticular relation‐ ships of D. punctatus. The median‐joining (MJ) haplotype networks were constructed to draw an unrooted network with Network 5 (Fluxus Technology Ltd.) based on COI gene, as well as ITS1 and ITS2. 2.4 | Population genetic analysis deposited in GenBank (accession numbers: KF366914–KF367149 for All analyses were executed separately for the COI gene and the ITS COI, KF367150–KF367271 for ITS1, and KF367272–KF367452 for fragments. We used DnaSP 5.10 (Librado & Rozas, 2009) to assess | 7483 LI et al. TA B L E 1 Locations of populations of Dendrolimus punctatus sampled, sample sizes (N), frequencies of mtDNA haplotypes per population, and estimates of haplotype diversity (h), and nucleotide diversity (π) for mitotypes Pop. Abbr. Locality Lat. (N) Long. (E) N Nh Haplotypes nos. h (SD) π in ‰ 1 GXLC Luchuan, Guangxi 22.19 110.16 6 4 H6 (2). H11 (2). H13(1). H18 (1) 2 0.867 (0.129) 1.93 2 GXNN Naning, Guangxi 22.49 108.21 12 4 H14(8). H15 (2). H16 (1). H17 (1) 3 0.561 (0.154) 1.62 3 GDDQ Deqing, Guangdong 23.15 111.77 8 5 H 11(3). H22(2). H23 (1). H24 (1). H25 (1) 3 0.857 (0.108) 2.37 4 YNWS Wenshan, Yunnan 23.22 104.15 6 4 H1 (1). H2 (3). H3 (1). H4 (1) 5 0.8 (0.172) 3.51 5 GXBS Baise, Guangxi 23.53 106.37 10 3 H46 (1). H47 (5). H48(4) 8 0.644 (0.101) 6.2 6 GDXN Xingning, Guangdong 24.15 115.73 4 2 H2(2). H7(2) 1 0.667 (0.204) 1.13 7 YNSL Shilin, Yunnan 24.45 103.16 6 2 H2 (4). H41 (2) 3 0.533 (0.172) 3.4 8 FJSH Shanghang, Fujian 25.02 116.25 9 3 H38 (5). H39 (1). H40 (3) 2 0.639 (0.126) 1.23 S 9 GZXY Xingyi, Guizhou 25.05 104.54 5 3 H 6(1). H19 (3). H20 (1) 12 0.7 (0.218) 8.16 10 FJWP Wuping, Fujian 25.10 116.10 11 6 H11 (3). H18 (2). H22 (3). H35 (1). H36 (1). H37 (1) 4 0.873 (0.071) 2.7 11 GXQZ Quanzhou, Guangxi 25.55 111.04 20 8 H5 (1). H6 (2). H8 (8). H9 (2). H10 (1). H11 (4). H12 (1). H13 (1) 7 0.811 (0.071) 4.87 12 YNYR Yongren, Yunnan 26.03 101.40 12 4 H4 (3). H5 (1). H6 (7). H7 (1) 3 0.636 (0.128) 1.49 13 GZGY Guiyang, Guizhou 26.38 106.37 8 4 H5 (1). H6 (3). H18 (1). H19 (3) 11 0.786 (0.113) 102 14 GZHZ Hezhang, Guizhou 27.07 104.43 7 3 H6 (3). H11 (3). H21 (1) 3 0.714 (0.127) 1.94 15 JXYC Yichun, Jiangxi 27.47 114.23 12 5 H5 (1). H6 (1). H22 (4). H23 (4). H45 (2) 5 0.803 (0.078) 3.27 16 JXGA Gaoan, Jiangxi 28.25 115.22 11 6 H2(4). H3(1). H41(2). H42(2). H43(1). H44(1) 5 0.855 (0.085) 3.83 17 ZJJS Jiangshan, Zhejiang 28.44 118.37 10 4 H27(3). H28(2). H29(4). H30(1). 6 0.778 (0.091) 4.72 18 ZJQX Quxian, Zhejiang 29.03 119.11 11 4 H1(1). H2(6). H3(3). H26(1) 3 0.673 (0.123) 1.48 19 ZJLX Lanxi, Zhejiang 29.12 119.28 9 6 H11(2). H18(1). H31(3). H32(1). H33(1). H34(1) 8 0.889 (0.091) 6.8 20 SCLX Luxian, Sichuan 29.15 105.38 4 1 H2(4) 0 0 0 21 SCHY Huaying, Sichuan 30.38 106.77 8 5 H6(3). H9(1). H22(2). H23(1). H37(1) 5 0.857 (0.108) 3.46 22 AHQS Qianshan, Anhui 30.57 116.34 12 3 H2(10). H3(1). H46(1) 2 0.318 (0.164) 0.57 23 HBCD Chengde, Hebei 40.59 118.53 35 3 H49(15). H50(8). H51(12) 2 0.666 (0.032) 1.65 Note: Voucher specimens are deposited in the Herbarium of Capital Normal University (CNU). population genetic diversity, including the number of sequences (N), number of haplotypes (Nh), haplotype diversity (h), and nucleotide 2.5 | Demographic history analysis diversity (π). The program PERMUT (available at www.pierr​oton. To evaluate the population history of D. punctatus, we calculated inra.fr/genet​ics/labo/Softw​are/) was applied to estimate the aver‐ Tajima's D (1989) and Fu's Fs (1997) neutrality test statistics age gene diversity within a population (HS), total gene diversity (HT ), based on the mitochondrial DNA dataset and calculated the sig‐ and between‐population divergence (GST, NST ) with 1,000 permuta‐ nificance with 10,000 simulations (Rogers & Harpending, 1992; tions tests. Schneider & Excoffier, 1999). Mismatch distribution analysis Analysis of molecular variance (AMOVA), the average values of was also conducted to detect the population expansion of the population differentiation (FST ), and Mantel test were all calculated D. punctatus with Arlequin V3.5 software (Excoffier & Lischer, to detect the population structure with Arlequin V3.5 software 2010); moreover, 1,000 parametric bootstrap replicates were (Excoffier & Lischer, 2010). The significance was tested using 10,000 used to test the suitability of observed mismatch distributions permutations. The spatial genetic pattern was examined by spatial to the theoretical distribution under a sudden expansion model analysis of molecular variance (SAMOVA) using SAMOVA V1.0 (Rogers & Harpending, 1992). When a prominent expansion event (Dupanloup, Schneider, & Excoffier, 2002). was identified, the parameter value for the mode of the mismatch 7484 | LI et al. distribution (τ) was applied to estimate time since expansion (in was evaluated by both the area under the curve (AUC) of the re‐ generations) using the equation t = τ/2u (Rogers, 1995). In this ceiver operating characteristic (ROC) plot and the binary omission study, u was calculated as u = μ × k × g, where μ represents the rate over ten replicate runs. Area under the curve is a compos‐ number of substitutions per site per year (s s −1 −1 y ), k represents ite measure of model performance, which can equally weigh the the average DNA sequence length, and g represents the genera‐ omission error and commission error, and has an index of suitability tion time in years. between 0 and 1. The value of AUC that is >0.5 indicates the pre‐ Bayesian skyline plots (BSPs) implemented in BEAST v2.0 dicted results of the model are better than those predicted sto‐ were also used to evaluate the timing of the population expansion chastically, and an AUC > 0.9 usually indicates excellent predictive (Bouckaert et al., 2014). For both COI and ITS fragments, the pro‐ power (Ye, Zhu, Chen, Zhang, & Bu, 2014). gram was run with three independent runs steps (30 million simu‐ lations), and we discarded the burn‐in of the first 20% steps. The evolutionary models using AIC were selected for COI (GTR) and ITS1 (TVMef+I) using jModeltest2 (Darriba et al., 2012). The pro‐ 2.7 | Populations identifying with distance‐based barcoding methods posed conventional mutation rates for insect mitochondrial COI The distance between intraspecific variation (the DNA barcoding gene 2.3% per million years were used to estimate coalescent time gap) is considered an important criterion in DNA barcoding. To ex‐ (Brower, 1994). After multiple runs, similar results and convergence plore the feasibility of populations that identify within D. punctatus, to stationarity were reached with effective sample size (ESS) of at DNA barcoding gap analysis was performed based on distance‐ and least 400. The Bayesian skyline plot (BSP) was constructed for each character‐based methods (Dai et al., 2012). The “best close match” data set using the program TreeAnnotator included in BEAST v 2.0 (BCM) (Meier, Shiyang, Vaidya, & Ng, 2006) and the minimum dis‐ (Bouckaert et al., 2014). tance (MD) (Zhang et al., 2012) are two distance‐based methods that proved to be effective in species identification in DNA barcod‐ 2.6 | Ecological niche modeling ing studies. To examine whether the correct local populations could be identified with these methods, we therefore performed the two The ecological niche model was a useful approach to visualize protocols with “single‐sequence‐omission” or “leave‐one‐out” simu‐ the population demographic history of D. punctatus during the lations. Here, we removed one sequence at a time and used it as a Quaternary, especially LGM. Climatic data from much earlier gla‐ query, with all other sequences remaining as the reference database. ciation could not be obtained; hence, we used present and LGM cli‐ We performed 500 random replications for each dataset. Moreover, matic data. If the distribution pattern of D. punctatus was related several species of Dendrolimus, D. superans, D. kikuchii, and D. houi to historical climatic changes, the distribution range would be ex‐ were chosen for interspecific level identification. We fully investi‐ pected to have obvious contracted during the LGM period (Carstens gated success rates of barcoding identification at both interspecific & Richards, 2007). and intraspecific levels with each single barcode, with their corre‐ A total of 107 occurrence records were collected for niche sponding 95% confidence intervals (CIs) computed using equation modeling analysis, including 23 records from the specimens in this (20) from Zhang et al. (2012). The BCM and MD were performed study and records from published documents. Bioclimatic variables following the protocol of the program TaxonDNA and package MD were downloaded from the WordClim website (V1.4; http://www. (Meier et al., 2006; Zhang et al., 2012). world​clim.org/). We compared the suitable habitats and potential distributions of D. punctatus in two periods: the present day and the LGM. The environmental layer of the present and LGM were built at a resolution of 2.5 min, and the LGM dataset was based on the Community Climate System Model (CCSM). A distribution 3 | R E S U LT S 3.1 | Genetic variation analyses was generated using 15 climatic variables from the WorldClim da‐ A total of 45 polymorphic sites were detected in the COI dataset, tabase (all bioclimatic variables except for BIO7, 8, 17, and 18) for which contained 43 parsimony informative sites. These polymor‐ the current climate, with an estimate based on other distribution phic sites defined 51 unique haplotypes (H1–H51) from the 236 localities that were collected (Hijmans, Cameron, Parra, Jones, & sampled individuals of D. punctatus, 18 of which were found in single Jarvis, 2005). The ecological niche model was constructed using individuals. A quarter of haplotypes (15 of 51) were unique to single the maximum entropy machine learning algorithm implemented in populations, while others were shared (Table 1). H2 was the most MAXENT (V3.3.3k; Phillips, Anderson, & Schapire, 2006), which abundant haplotype shared by 33 individuals, mainly distributed in proved to perform well compared with other methods, especially the southwest region of China, followed by H6, which also occurring those with relatively smaller sample sizes. Analysis with default in six populations. Overall nucleotide diversity was π = 0.0075 and program settings (cumulative output, convergence threshold (1 of varied across populations (ranging from 0.0011 to 0.0082). Within iterations (500)) was executed in ten replicates with default set‐ all populations, GZXY (see population abbreviation in Table1) had tings; 70% of the sites were set to train the model, and 30% of the the highest π value (Table 1). Haplotype diversity was h = 0.952 sites were set to test the model predictions. Model performance (ranging from 0.000 to 0.889) at the species level, and ZJLX (see | 7485 LI et al. TA B L E 2 Estimate of nucleotide diversity (Pi), haplotype diversity (Hd), total gene diversity (HT ), average gene diversity within populations (HS), interpopulation differentiation (GST ), and the number of substitution types (NST ) (mean + SE in parentheses) as indicators of mtDNA haplotypes and ITS ribotypes diversity Pi Hd HT HS GST NST Mitochondrial DNA Dendrolimus punctatus 0.00751 0.952 0.947 (±0.0204) 0.692 (±0.0424) 0.269 (±0.0383) 0.580 (±0.0486)** South Region 0.00858 0.912 0.916 (±0.0465) 0.580 (±0.1001) 0.367 (±0.0945) 0.587 (±0.0829)* Southeast Region 0.00643 0.942 0.960 (±0.0203) 0.743 (±0.0440) 0.226 (±0.0415) 0.527 (±0.0597)** Internal Transcribed Spacer 1 (ITS1) Dendrolimus punctatus 0.00838 0.773 0.743 (±0.0649) 0.595 (±0.0596) 0.199 (±0.0635) 0.373 (±0.0783)** South Region 0.00853 0.900 0.550 (±0.0808) 0.547 (±0.0933) 0.006 (±0.0447) 0.086 (±0.0382)** Southeast Region 0.00688 0.630 0.916 (±0.0308) 0.735 (±0.0638) 0.197 (±0.0756) 0.457 (±0.0960)** Internal Transcribed Spacer 2 (ITS2) Dendrolimus punctatus 0.00277 0.818 0.840 (±0.0350) 0.587 (±0.0496) 0.301 (±0.0633) 0.331 (±0.0763) NS South Region 0.00284 0.822 0.882 (±0.0453) 0.517 (±0.0915) 0.414 (±0.1015) 0.440 (±0.1027)NS Southeast Region 0.00268 0.795 0.801 (±0.0526) 0.666 (±0.0571) 0.169 (±0.0775) 0.165 (±0.0789)NS Abbreviation: NS, not significant. *p < 0.05, significant. **NST is significantly different from GST (p < 0.01). NS p > 0.05, not significant. population abbreviation in Table1) had the highest h value (Table 1). the analysis of mitochondrial sequences, the topology of phylogenetic Within‐population gene diversity (HS) was much lower than total trees based on two methods were similar (Figure 3a). The haplotypes gene diversity (HT ) (0.694 and 0.942, respectively, Table 2). of D. punctatus were split into several discrete clades. The monophyly The length of aligned ITS1 sequences was 670 bp including 69 of D. punctatus was well supported. These 51 haplotypes were clus‐ polymorphic sites. Thirty‐eight different ITS1 haplotypes (ribo‐ tered into three multiple haplotype clades (Clades I, II, and III) and types1) were recovered (R1–R38), two of which occurred in sin‐ two monotypic clades (Clades IV and V) (Figure 3a). The haplotype gle populations separately. Meanwhile, the other haplotypes were network (Figure 3b) displayed a more detailed relationship among shared by more than two populations. The most widespread ribo‐ 51 haplotypes. The populations located in the southwest region har‐ types1 was R26 occupied by 19 populations. Ribotype1 diversity bored four haplotype clusters, and only two populations possessed was estimated to be h = 0.773 at the species level, ranging from single haplotype clusters. The haplotypes nested in Clades I and II 0.000 to 1.000 in different populations. At the species level, were mostly found in the eastern part of China (Figure 2). nucleotide diversity was 0.008 (p = 0.022), but it varied across For the ITS1 dataset, the topological structure of the phyloge‐ populations, ranging from 0.000 to 0.0189, with the GDXN (see netic trees, reconstructed by two methods, were similar (Figure 4a). population abbreviation in Table1) population having the highest All 38 ribotypes were divided into two clades, in which there were nucleotide diversity. Similar results were detected in the ITS2 se‐ 23 ribotypes1 formed a well‐support clade (Clade I). The ribotype1 quences dataset including 22 polymorphic sites in 181 individuals. network (Figure 4b) reflected similar relationships with more details, A total of 29 different ITS2 haplotypes (ribotypes2) were defined and the obvious star‐like structure network was generally inter‐ (N1–N29). Overall nucleotide diversity was 0.0028, and haplotype preted as a sign that the population underwent population expan‐ diversity was 0.818 (ranging from 0.167 to 1.000). N15 was the sion. R26 has the highest haplotype frequency (19 of 23). Results most popular ribotypes2, which was found in 16 populations of the based on ITS2 fragment analysis agreed with ITS1. However, due to southern region. Similar results were detected comparing with the the relatively few polymorphic sites, the distinct topological struc‐ mitochondrial dataset. Within‐population gene diversity (HS) was ture was not detected based on both ML and BI phylogenetic trees lower than total gene diversity (H T ) (0.156 and 0.879, respectively; (Figure 5a). Future MJ network tests showed that N15 was in the Table 2); however, only two populations harbored single ribotype center position, which is in line with the wide distribution of N15 in for the ITS1 dataset, and only one population for ITS2 sequences. 18 out of 23 populations (Figure 5b). 3.2 | Phylogenetic relationships and lineages divergence 3.3 | Population structure analysis Phylogenetic analyses were performed using the ML and BI methods. ples based on both mitochondrial and nuclear sequences (Table 1). The GTR+G model was selected by COI and ITS datasets. Based on Further examinations using permutation tests showed a prominent Significant genetic differentiation was identified in D. punctatus sam‐ 7486 | LI et al. F I G U R E 3 Phylogenetic relationships obtained by analysis of mtDNA haplotypes. (a) ML tree with numbers above the branches indicating bootstrap values greater than 50% for ML analysis. (b) NETWORK‐derived genealogical relationship. The sizes of the circles in the network are proportional to the observed frequencies of the haplotypes. The small black bars represent mutation steps, and the gray dots represent missing haplotype. For both subfigures, different colors represent different haplotype clusters higher value of NST than GST (p < 0.01), indicating a strong phylo‐ values of K (FCT values ranged from 0.217 to 0.344 for COI, from geographic structure in D. punctatus. The Spatial AMOVA (SAMOVA) 0.313 to 0.447 for ITS1, and from 0.193 to 0.499 for ITS2); however, method represents a useful method to define partitions of sampling the first level of divergence (K = 2) revealed a well‐defined group sites that are maximally differentiated from each other without any including those populations from the east region of China with a priori assumption of population structure. However, the analysis high frequency of mitochondrial haplotype Clade I, which were dif‐ did not clearly reveal one single group of maximally differentiated ferentiated from the remaining populations. The composition of the populations, as FCT values increased progressively with increasing groups detected by SAMOVA that had increasingly larger values of K LI et al. | 7487 F I G U R E 4 Phylogenetic relationships obtained by analysis of ITS1 ribotypes. (a) ML tree with numbers above the branches indicating bootstrap values greater than 50%. (b) NETWORK‐derived genealogical relationship F I G U R E 5 Phylogenetic relationships obtained by analysis of ITS2 ribotypes. (a) ML tree with numbers above the branches indicating bootstrap values greater than 50%. (b) NETWORK‐derived genealogical relationship 7488 | LI et al. mainly corresponded to the geographical distribution of haplotypes significantly negative (Table 3), which indicated that the population (Figure 2). Consistent results of two distinguished groups were con‐ of D. punctatus underwent rapid expansion. The estimated time of firmed by ITS1 sequence analyses. Thus, we divided the 23 popula‐ this expansion was 0.232 Mya (Table 3). The population expan‐ tions collected into two geographic groups: one for the eastern of sion of haplotype Clade I and Clade II were also detected based on China, and the other for the southwest of China. Pairwise FST was all mismatch distribution analysis, as well as the Bayesian population significant (p < 0.01) in 1,000 permutation tests between two geo‐ dynamic analysis (Table 3). The results obtained from the Bayesian graphic regions: 0.393 for COI, 0.363 for ITS1, and 0.523 for ITS2, population dynamic analysis for all individuals are shown in Figure 7, respectively. This suggested high genetic differentiation between which indicates that the population experienced conspicuous ex‐ these two geographic groups. pansion, in addition to one obvious bottleneck that occurred about 0.02 Mya. For the ITS loci, recent population expansion was inferred 3.4 | Demographic history of D. punctatus As previously mentioned, a star‐like structure in a network analy‐ sis is generally interpreted as a sign of population expansion/rapid at 0.025 Mya (ITS1) and 0.17 Mya (ITS2) (Figure 7). 3.5 | Species distribution modeling growth (Rogers & Harpending, 1992). The haplotype network results For this study, ecological niche modeling was used to explore the re‐ indicated historical population expansion in the masson pine moth lationships between genetic diversity and possible glacial refugia. A history. This finding was further supported by the mismatch distri‐ species distribution model for D. punctatus was built and the average bution analysis. Based on COI, ITS1, ITS2, or combined ITS datasets, AUC value was 0.917, which is considered to correspond to a useful the null hypothesis that showed sudden population expansion his‐ predictive model. Under the present climatic conditions, the model torically existed could not be rejected (p = 0.427 for COI; p = 0.049 revealed high climatic habitat suitability for D. punctatus across the for ITS1; p = 0.409 for ITS2; p = 0.752 for the combined ITS1 and ITS2; south region of China. For the predicted suitable distribution, there Figure 6). However, the raggedness indices and observed variance was a significant difference between the present and LGM periods. (SSD) were not significantly different from the expectation that sug‐ The distribution area was contracted to the southern and southeast gested strong and wide range gene flow occurred in the population parts of China in LGM; therefore, the two lineages were isolated in demographic history. In addition, Tajima's D and Fu's Fs were also the glacial periods (Figure 8). Interestingly, the results also revealed FIGURE 6 Mismatch distribution analysis based on (a) COI, (b) ITS1, (c) ITS2, and (d) the combination of ITS1 and ITS2 | 7489 LI et al. TA B L E 3 Values for τ, SSD, and RAG (with probability p values) from mismatch analysis of mitotype variation, values for Tajima's D and Fu's FS (with p values) Mismatch distribution Groups τ(t) t1–t2 (Mya) Dendrolimus punctatus 4.826 0.232–0.358 Western region 2.351 Neutrality tests RAG (p value) Tajima'D (p value) Fu's FS (p value) 0.003 (0.493) 0.006 (0.665) −0.180 (0.561) −17.827 (0.001)** – 0.014 (0.381) 0.024 (0.538) 0.686 (0.789) −4.083 (0.100) NS SSD (p value) Eastern region 5.078 0.244–0.375 0.004 (0.346) 0.015 (0.572) −0.476 (0.739) −25.080 (0.000)** Clade I 3.262 0.157–0.242 0.001 (0.907) 0.012 (0.962) −0.281 (0.459) −6.351 (0.024)* Clade II 1.117 0.057–0.083 0.004 (0.537) 0.062 (0.117) 0.203 (0.108) −6.620 (0.001)** Clade III 1.839 – 0.006 (0.712) 0.051 (0.151) 1.123 (0.217) −1.619 (0.076) NS Clade V 0.286 – 0.0004 (0.214) 0.265 (0.345) −1.006 (0.201) −0.095 (0.369) NS Note: Related value of Clade IV was not calculated due to insufficient individuals (2 sequences). Abbreviations: g, estimate of population exponential growth rate; NS, not significant; RAG, the Harpending's raggedness index; SSD, sum of squared deviations; t1, time in million years derived from 0.0115 s s−1 y−1; t2, time in million years derived from 0.0177 × 10 −6 s s−1 y−1; τ, time in number of generations elapsed since a sudden expansion episode. *p < 0.05. **p < 0.01. NS p > 0.05. that the southern and southwest regions in China may have served success rate of 25.20% (95%CI: 21.59%–29.18%) and 13.20% (95%CI: as feasible refugia during the maximum of the Pleistocene glacia‐ 10.51%–16.44%), respectively. The MD method (Figure 9d and f) re‐ tions. These findings further support the conclusions drawn by pop‐ flected the difficulties in local population DNA barcoding. ulation demographic history analysis. 3.6 | DNA barcoding in population identification In general, a large DNA barcoding gap makes species/population dis‐ crimination possible and simple. Conversely, small or negative DNA 4 | DISCUSSION 4.1 | Genetic diversity and population structure in D. punctatus barcoding gaps blur species/population boundaries and hamper Our survey of mtDNA variation of D. punctatus detected several species/population assignation in DNA barcoding. The COI barcode phylogeographical groups within this hazardous insect across the showed an average between‐population K2P distance of 0.013, which south region of China. This finding suggests that in the past, the dis‐ was about 1.78 times larger than the mean within‐population dis‐ tribution of this species was fragmented into several isolated refugia, tance. However, no positive DNA barcoding gap for the COI barcode where the genetic divergence of mtDNA took place. However, due (Figure 9a) was detected, indicating the difficulty of distinguishing to the relative limited informative sites, no significant phylogeo‐ subpopulations of this species. Both ITS1 and ITS2 genes presented graphical pattern was revealed for ITS2 dataset. slightly larger between‐population genetic variation (0.0085 for ITS1; The mtDNA diversity was relatively high (HT = 0.947) in D. punc‐ 0.0029 for ITS2) compared with genetic variation within popula‐ tatus, which might reflect a longer evolutionary history and the wide tion (0.0077 for ITS1; 0.0018 for ITS2). The former was roughly 1.1 distribution of this species. However, the AMOVA analysis based on (1.1038) and 1.6 (1.6111) times larger than the latter. Similarly, there COI dataset revealed genetic differentiation among all 23 popula‐ were no positive barcoding gaps or a blurred discriminations among tions and most variation was from within populations (47.75%), fol‐ subpopulations with ITS1 and ITS2 (Figure 9c and e). lowed by variation among groups (28.74%) (Table 4). Consequently, Significant differences in success rates between interspecific and most of the species’ diversity was due to population divergence in intraspecific barcoding identification were detected with MD and BCM which distinct phylogeographical structure signals were detected methods (Figure 9b, d, and f). Intraspecific success rates were signifi‐ (NST = 0.580 > GST = 0.269, p < 0.01). cantly smaller than interspecific ones. For instance, based on the COI Further tests carried out with ITS1 were corroborated with gene, mean intraspecific success rate was 51.20% with a 95% confidence our COI findings, and distinct phylogeographic signals (NST > GST, interval (CI) of 46.83%–55.55% (MD), 35.40% (95% CI: 31.33%–39.68%) p < 0.01) were found in entire distribution regions, including west‐ (BCM), while corresponding interspecific success rates achieved values ern and eastern regions. A relatively low value of total diversity of 97.40% (95% CI: 95.60%–98.47%) and 96.60% (95% CI: 94.62%– was recorded for ITS dataset in D. punctatus (ITS1: HT = 0.743, ITS2: 97.87%), respectively (Figure 9b). Significantly smaller success rates HT = 0.840), which was lower than the mitochondrial dataset, largely were detected for nuclear markers ITS1 and ITS2, which obtained a low due to reduced genetic divergence (ITS1: GST = 0.199, NST = 0.373; 7490 | LI et al. F I G U R E 7 Population dynamic analysis of Dendrolimus punctatus Bayesian skyline plots for (a) COI gene and (b) ITS loci ITS2: GST = 0.331, NST = 0.301), as within‐population diversity was The AMOVA analysis of the mtDNA data revealed that within‐ also decreased compared with mtDNA (HS = 0.595 for ITS1 and population variation was greater than among‐region variation. HS = 0.587 for ITS2). The primary cause of this discrepancy is due Nevertheless, D. punctatus is split into two major genetic lineages to the different modes of inheritance, maternal for mtDNA and bi‐ based on the analysis of mtDNA and ITS1 (FCT = 0.127 and 0.313, parental for ITS. The latter method leads to higher levels of gene separately) (Table 4). This phenomenon could be explained as re‐ flow. Compared with the protein‐coding COI region, the noncoding peated population expansion and contraction caused by climate ITS marker was more sensitive at detecting gene flow among inter‐ oscillations during the ice age, which resulted in lineage differ‐ breeding populations. Similar inconsistencies were shared in many entiation and limited gene flow. Physical barriers and ecological other Lepidopteran insets, for example, Biston suppressaria (Cheng, climate factors usually played a key role in obstructing gene flow Jiang, Xue, et al., 2016; Cheng, Jiang, Yang, et al. 2016), Apocheima among populations (Fjeldså, Bowie, & Rahbek, 2012). Although cinerarius (Liu et al., 2016), Biston panterinaria (Cheng, Jiang, Xue, subtropical China has never been covered by extensive ice sheets et al., 2016; Cheng, Jiang, Yang, et al., 2016), and Hyles euphorbiae during the ice age, iterative alternation of glacial and interglacial (Mende & Hundsdoerfer, 2013). periods particularly caused distribution changes of both plants | 7491 LI et al. the northwestern and eastern lineages was likely to be affected by these ridges. These mountains were important faunal dividing criterions in subtropical China. Similar geographical separations have been reported in many pervious researches, including the Leucodioptron canorum (Li et al., 2009), Squalidus argentatus (Yang et al., 2012), and Microvelia horvathi (Ye et al., 2014). Meanwhile, present limited gene flow between southwest and east regions was also caused by the decreased ability of long‐dis‐ tance dispersal by D. punctatus (Hou, 1987). As a typical phytoph‐ agous insect with intermediate mobility, D. punctatus exhibited a steep genetic decline with distance. The populations around the dis‐ tributional center tended to have higher levels of genetic variations than the peripheral ones (Table 1; Figure 2). High genetic diversity in the center of the range of distribution reflected that the gene flow was strong among closely situated populations, which apparently was due to its ability to disperse in nearby areas. Geographic barri‐ ers such as Mt Wu, Mt Xuefeng, and Yungui Plateau (YGP) weakened the population diffusion and gene flow of D. punctatus, especially between the southwest and east regions. 4.2 | Pleistocene range expansion and potential refugia of D. punctatus High polymorphism detected in the D. punctatus with mtDNA and ITS may reflect the wide distribution of this forest pest in south‐ ern China, with large and relatively stable population sizes. This was compared with other pine pests distributed in Europe (e.g., Dendrolimus sibiricus, Dendrolimus pini, and Dendrolimus sibiricus), which experienced drastic reduction in population size due to the coverage of ice sheet during the Quaternary (Kononov et al., 2016; Mikkola & Ståhls, 2008). Both our phylogenetic and network analyses showed that D. punctatus have experienced population expansion in subtropical F I G U R E 8 Ecological niche modeling for Dendrolimus punctatus during the present day (a) and the LGM (b). Different colors correspond to different fitting indices with low in red and high in green China. According to mismatch distribution and Bayesian popula‐ tion dynamic analysis based on the mtDNA dataset, the popula‐ tion expansion started at ca. 0.23 Mya. Cui et al. (2011) reported a detailed partition criterion of Quaternary glaciations in China and concluded that six major glaciations occurred since Wangkun and animals (Liu, 1988; Qiu et al., 2011). Continuous and intensive Glaciation (0.7–0.5 Mya). The expansion of D. punctatus initi‐ tectonic activities had already begun in subtropical China during ated at the interglacial period of Guxiang Glaciation (Penultimate the Miocene and Pliocene (Webb & Bartlein, 1992), which made a Glaciation) (also known as MIS7). During the interglacial periods great effort in shaping the distribution and differentiation of local of Penultimate Glaciation, populations could be effectively pre‐ organism. The northeast–southwest trending mountain ridges served in the mountainous areas of southern China. In addition, in subtropical China, such as Mt Wu, Mt Xuefeng, and Mt Wuyi, D. punctatus is one of the most serious and economically damaging acted as barriers against the inflow of cold air from the north‐ insect pests, which have the genetic characteristics of population west and maintained warm moist air coming from the Pacific, re‐ outbreak initiated by high temperature and drought (Zhang et al., sulting in a humid climate in the eastern region and relative arid 2004). The interglacial periods provide feasible climate conditions environment in the southwest. Moreover, the Qinghai‐Tibetan for the population outbreak and allow the maintenance of rela‐ Plateau (QTP) and Mt Hengduan obstructed the dampness from tively large population sizes, thus retaining high genetic diversity. the Indian Ocean, which further exacerbated the drought in this Interestingly, the event of demographic bottlenecks was detected region. Mt Xuefeng also acted as a protective barrier against dif‐ based on the Bayesian skyline plot analysis of mtDNA dataset. This ferent climate environments on the eastern edge of the Yungui sudden population shrink occurred at ca. 0.02 Mya, which belong Plateau (YGP) (Holt et al., 2013). Restricted gene flow between to the LGM period. 7492 | LI et al. F I G U R E 9 Interspecific and intraspecific DNA barcoding success rates of Dendrolimus punctatus and related species. (a) Intraspecific barcoding gap analysis based on COI gene. Blue and red histograms represent distributions of between‐ and within‐population genetic variations (K2P distance), respectively, with red fitted normal distribution curve each. (b) Interspecific and intraspecific barcoding success rates based on COI barcode with MD and BCM methods; (c) Intraspecific barcoding gap analysis based on ITS1 gene; (d) Interspecific and intraspecific barcoding success rates based on ITS1 with MD and BCM methods. (e) Intraspecific barcoding gap analysis based on ITS2 gene; (f) Interspecific and intraspecific barcoding success rates based on ITS2 with MD and BCM methods | 7493 LI et al. 1.803 303.188 180 4.152 464.516 121 3.757 861.394 235 Total Abbreviations: df, degrees of freedom; FCT, correlation within groups relative to total.; FSC , correlation within populations relative to group; FST, correlation within populations relative to total; NS, not significant; PV, percentage of variation; SS, sum of squares; VC, variance components. *p < 0.05, significant. **NST is significantly different from GST (p < 0.01). NS p > 0.05, not significant. FCT = 0.193NS 35.64 0.599 105.724 158 FCT = 0.313* 43.72 1.645 201.902 99 FCT = 0.217* 47.75 1.782 386.187 213 Within populations FST = 0.524** 37.02 0.656 21 21 220.384 0.956 23.51 FST = 0.554** 21 126.465 1.238 25.97 FST = 0.363** Among regions Among populations 30.31 1.269 136.149 1 1 254.823 1.019 28.74 FSC = 0.565** df Fixation index PV (%) VC 107.660 FSC = 0.486** 27.34 0.548 89.804 1 FSC = 0.282** PV (%) VC SS df SS VC PV (%) Fixation index ITS2 ITS1 Mitochondrial DNA SS df Source of variation TA B L E 4 Analysis of molecular variance (AMOVA) conducted on mtDNA haplotypes variation and ITS ribotypes for within and among populations of Dendrophilus punctatu Fixation index Dendrolimus punctatus is a typical phytophagous insect, in which its patterns of populational distribution and dynamics have significant correlation not only with the meteorological factors but also its host plants. Although the main host plant of D. punctatus is the masson pine, P. massoniana and its caterpillars have been reported to feed on other conifers, such as Pinus thunbergii, Pinus kwangtungensis, and Pinus taeda (Cai, 1995; Hou, 1987). The exten‐ sive hilly areas of subtropical China (especially between 22°N and 34°N) provide advantages to harbor temperate deciduous forests and coniferous (Yu et al., 2000). Many researchers have argued that lots of local plant species experienced long‐term fragmenta‐ tion and refugial survival, such as Cathaya argyrophylla (Pinaceae) and Eurycorymbus cavaleriei (Sapindaceae), and several postulating separate refugia (such as Mts Dayao, East Yungui Plateau, and Mts Nanling) in subtropical China (Qiu et al., 2011; Wang, Gao, Kang, Lowe, & Huang, 2009). In addition, clear phylogeographic histories of P. kwangtungensis (Pinaceae), a closely related species of the host plant P. massoniana, were demonstrated among group FST differen‐ tiation, suggesting that Yungui Plateau and the Mts Nanling were a major refugium (Tian, López‐Pujol, Wang, Ge, & Zhang, 2010). Furthermore, more recent studies have confirmed that P. masso‐ niana also existed in multiple glacial refugia fostering populations with high genetic diversity (Ge et al., 2012; Zhou et al., 2010). The high genetic variation of D. punctatus in the southwest region is an alternative explanation for the retention of ancestral polymor‐ phisms. A similar genetic divergence pattern was also detected in our study. The populations form Yungui plateau (GZXY and GZHZ, see population abbreviation in Table1) and Mts Guangzhou Province (FJSH, see population abbreviation in Table1) harbored haplotypes from separate haplotype clusters, suggesting a shared semblable refugia with its host plants. Temperate evergreen forest, such as P. massoniana, recolonized the hilly regions of south China because of the increasing tempera‐ ture since the Holocene (Harrison et al., 2001). Increased number of host plants provided great opportunity for sustaining population expansion of D. punctatus. In our study, a noteworthy trend revealed higher diversity in the southern region than in the east and north regions, reflecting a common scenario of postglacial, northward re‐ colonization. Two stages of population growth are also well consis‐ tent with the last interglacial high sea level (MIS5, 0.13–0.07 Mya) and postglacial‐Holocene warm climate (0.018–0 Mya). The warm climate and high sea level influenced interglacial continental vege‐ tation distribution. The extinction of the populations in the northern parts of China, caused by migration to the south during the glacial maximum, might have survived in separate refugia. Our survey of mtDNA variation throughout the geographical distribution of D. punctatus suggests that several isolated refugia were existed where mtDNA divergence occurred. In addition, population dynamic analysis indicated that the D. punctatus species experienced population expansion, and an obvious northward recolonization was detected. The haplotype distribution of the present investigation seems further to indicate that the common fixture of the same haplotype among the disjunct 7494 | LI et al. populations was mainly caused by the founder effect during the a phylogeographical hypothesis for further testing; however, given population recolonization in this region. The highest latitude popu‐ the small sample size and limited molecular markers, many low‐copy lation (pop 23, HBCD, see population abbreviation in Table1) has rel‐ nuclear genes and mitochondrial genomes are required to further atively low genetic divergence even though much more individuals examine the genomic divergence within D. punctatus. were used, which indicates the founder effect may be an important factor that influences genetic differentiation of D. punctatus. 4.3 | Applicability of DNA barcoding at population level AC K N OW L E D G M E N T S This research was supported by China National Funds for Distinguished Young Scientists (Grant No. 31425023), by Natural Science Foundation of China (Grant Nos. 31601877, 31772501 and To assign unknown individuals to certain populations within a spe‐ 31400191), by Beijing Municipal Natural Science Foundation (Grant cies or to identify population from DNA sequences is not a new ap‐ No. 5172005). proach (Lou & Golding, 2010). Here, we present a study in which we systematically investigated the population structure and dynamic of D. punctatus with a variety of DNA barcoding methods, based on both commonly used COI gene, and two alternatives genes, ITS1 and CONFLIC T OF INTEREST None declared. ITS2. The intraspecific success rate of barcoding identification was found to be significantly lower than the interspecific ones, indicat‐ ing the difficulty in barcoding populations (Figure 9). For the COI gene, among 23 populations of D. punctatus, only one single local population (pop23, HBCD) formed a monophyletic clade (Figure 3). Moreover, compared with the COI gene, less population structure was also found for ITS genes on the phylogenetic trees (Figure 4,5). The results may not be surprising since the coalescence time is gener‐ AU T H O R C O N T R I B U T I O N S Jing Li wrote the MS and Qian Jin performed the experiment. Jing Li and Qian Jin analyzed all the data together. Geng‐ping Zhu and Chong Jiang provided great help in ecological niche modeling analy‐ sis. Ai‐bing Zhang defined the project as the main supervisor and considered the whole analysis. ally shorter than the time of different species to their most common ancestor (TMCA), therefore, no large enough variation or mutations accumulated during this relatively short time period. Furthermore, The AMOVA analysis of the mtDNA data suggested that strong gene flow existed in the D. punctatus populations, which also impaired the success rate of population assignments and resulted in no positive barcoding gaps being detected (Figure 9). DATA AC C E S S I B I L I T Y DNA sequences: KF367149 for GenBank COI, accession numbers: KF367150–KF367271 for KF366914– ITS1, and KF367272–KF367452 for ITS2. BOLD system accession numbers: GBGL14669–GBGL14904 for COI. Despite the fact that the sequence information collected from DNA barcoding was not sufficient to solve population‐level ques‐ tions, it reflected genetic diversity among different population (Bazin, ORCID Glémin, & Galtier, 2006; Moritz & Cicero, 2004). It should be noted, Geng‐ping Zhu https://orcid.org/0000-0001-6823-5840 however, that DNA barcoding markers have the valuable potential Ai‐bing Zhang https://orcid.org/0000-0003-3450-5421 to investigate the historical population dynamics of D. punctatus. It would be interesting to investigate further the applicability of bar‐ coding information in the study of genetic diversity of other species. 5 | CO N C LU S I O N OPEN RESEARCH BADGE S This article has been awarded an Open Data and Open Materials. All materials and data are publicly accessible via the Open Science Framework at https​://doi.org/10.5061/dryad.2df87g2. Learn more We investigated the population structure and dynamics of D. puncta‐ about the Open Practices badges from the Center for Open Science: tus based on the genetic variation of mtDNA as well as nuclear mark‐ https​://osf.io/tvyxz/​wiki. ers. A history of postglacial recolonization of the populations and multiple refugia during the ice age of the Quaternary was inferred. Many previous phylogeographic studies have paid close attention to the effects of very large mountainous systems such as the Mts Hengduan and Mts Qinling, but less attention was paid to the effect of the smaller mountain regions in the subtropical regions of China. The findings of this study proved that several potential refugia and geographic barriers existed in these hilly areas. These results provide REFERENCES Alexandros, S. (2006). RAxML‐VI‐HPC: Maximum likelihood‐based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics, 22(21), 2688–2690. Bazin, E., Glémin, S., & Galtier, N. (2006). Population size does not influ‐ ence mitochondrial genetic diversity in animals. Science, 312(5773), 570–572. LI et al. Bouckaert, R., Heled, J., Kühnert, D., Vaughan, T. G., Wu, C. H., Xie, D., … Drummond, A. J. (2014). BEAST2: A software platform for Bayesian evolutionary analysis. PLoS Computational Biology, 10(4), e1003537. Brower, A. V. Z. (1994). Rapid morphological radiation and convergence among races of the butterfly Heliconius erato inferred from patterns of mitochondrial DNA evolution. Proceedings of the National Academy of Sciences of the United States of America, 91, 6491–6495. Cai, X. M. (1995). Studies on dynamics of population of Dendrolimus punctatus. Journal of Zhejiang Forest Science and Technology, 15, 1–84. (in Chinese with English summary). Carstens, B., & Richards, C. L. (2007). Integrating coalescent and eco‐ logical niche modelling in comparative phylogeography. Evolution, 61, 1439–1454. Cheng, R., Jiang, N., Xue, D., Li, X., Ban, X., & Han, H. X. (2016). The evolutionary history of Biston suppressaria (Guenée). (Lepidoptera: Geometridae). related to complex topography and geological history. Systematic Entomology, 41(2), 732–743. Cheng, R., Jiang, N., Yang, X. S., Xue, D. Y., Liu, S. X., & Han, H. X. (2016). The influence of geological movements on the population differen‐ tiation of Biston panterinaria (Lepidoptera: Geometridae). Journal of Biogeography, 43, 691–702. Cui, Z. J., Chen, Y. X., Zhang, W., Zhou, S. Z., Zhao, L. P., Zhang, M., & Li, C. C. (2011). Research history, glacial chronology and origins of Quaternary glaciations in China. Quaternary Science, 31(5), 749–764. Dai, Q. Y., Gao, Q., Wu, C. S., Chesters, D., Zhu, C. D., & Zhang, A. B. (2012). Phylogenetic reconstruction and DNA barcoding for closely related pine moth species (Dendrolimus) in China with multiple gene markers. PLoS ONE, 7, e32544. Darriba, D., Taboada, G. L., Doallo, R., & Posada, D. (2012). jModelT‐ est 2: More models, new heuristics and parallel computing. Nature Methods, 9, 772. DeSalle, R. O. B. (2006). Species discovery versus species identifica‐ tion in DNA barcoding efforts: Response to Rubinoff. Conservation Biology, 20(5), 1545–1547. Dupanloup, I., Schneider, S., & Excoffier, L. (2002). A simulated annealing approach to define the genetic structure of populations. Molecular Ecology, 11, 2571–2581. Edgar, R. C. (2004). Muscle: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research, 32, 1792–1797. Excoffier, L., & Lischer, H. E. L. (2010). Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Molecular Ecology Resources, 10, 564–567. Fjeldså, J., Bowie, R. C. K., & Rahbek, C. (2012). The role of mountain ranges in the diversification of birds. Annual Review of Ecology, Evolution, and Systematics, 43(1), 249–265. Friberg, M., Schwind, C., & Roark, L. C. (2014). Floral scent contributes to interaction specificity in coevolving plants and their insect pollina‐ tors. Journal of Chemical Ecology, 40, 955–965. Friberg, M., Schwind, C., & Thompson, J. N. (2016). Divergence in selec‐ tion of host species and plant parts among populations of a phytoph‐ agous insect. Evolution and Ecology, 30, 723–737. Fu, Y. X. (1997). Statistical tests of neutrality of mutations against pop‐ ulation growth, hitchhiking and background selection. Genetics, 147, 915–925. Ge, X. J., Hsu, T. W., Hung, K. H., Lin, C. J., Huang, C. C., Huang, C. C., … Chiang, T. Y. (2012). Inferring multiple refugia and phylogeographical patterns in Pinus massoniana based on nucleotide sequence variation and DNA fingerprinting. PLoS ONE, 7, e43717. Hajibabaei, M., Singer, G., Hebert, P. D. N., & Hickey, D. (2007). DNA barcoding: How it complements taxonomy, molecular phylogenetics and population genetics. Trends in Genetics, 23, 167–172. Harrison, S. P., Yu, G., Takahara, H., & Prentice, I. C. (2001). Palaeovegetation: Diversity of temperate plants in East Asia. Nature, 413, 129–130. | 7495 Hebert, P. D. N., Cywinska, A., Ball, S. L., & deWaard, J. R. (2003). Biological identifications through DNA barcodes. Proceedings of the Royal Society B: Biological Sciences, 270, 313–321. Hewitt, G. M. (2000). The genetic legacy of the Quaternary ice ages. Nature, 405, 907–913. Hickerson, M. J., Carstens, B. C., Cavender‐Bares, J., Crandall, K. A., Graham, C. H., Johnson, J. B., … Yoder, A. D. (2010). Phylogeography's past, present and future: 10 years after Avise, 2000. Molecular Phylogenetics and Evolution, 54, 291–301. Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G., & Jarvis, A. (2005). Very high‐resolution interpolated climate surfaces for global land areas. International Journal of Climatology, 25, 1965–1978. Holt, B. G., Lessard, J. P., Borregaard, M. K., Fritz, S. A., Araújo, M. B., Dimitrov, D., … Rahbek, C. (2013). An update of Wallace’s zoogeo‐ graphic regions of the world. Science, 339, 74–78. Hou, T. Q. (1987). Lepidoptera: Lasiocampidae and Eupterotidae in in‐ sects of the three gorge reservoir area of Yangtze river. In X. K. Yang (Ed.). The pine caterpillars in China (pp. 1396–1406). Chong Qing, China: Chong Qing Press (in Chinese). Huemer, P., & Hebert, P. D. N. (2011). Cryptic diversity and morphology of high alpine Sattleria – A case study combining DNA barcodes and morphology (Lepidoptera: Gelechiidae). Zootaxa, 2981, 1–22. Jiang, W. B., He, H. Y., Li, Y. D., Ren, M. Y., Ma, Y. Z., Zheng, L. L., … Yu, W. D. (2016). Taxonomic status and molecular phylogeography of two sibling species of Polytremis (Lepidoptera: Hesperiidae). Scientific Reports, 6, 20820. Kononov, A., Ustyantsev, K., Wang, B. D., Mastro, V. C., Fet, V., Blinov, A., & Baranchikov, Y. (2016). Genetic diversity among eight Dendrolimus species in Eurasia (Lepidoptera: Lasiocampidae), in‐ ferred from mitochondrial COI and COII, and nuclear ITS2 mark‐ ers. BMC Genetics, 17(Suppl. 3), 157–166. https​://doi.org/10.1186/ s12863-016-0463-5 Kumar, S., Tamura, K., & Nei, M. (2004). Mega 3: Integrated software for molecular evolutionary genetics analysis and sequence alignment. Brief Bioinformatics, 5, 150–163. Lei, F. M., Qu, Y. H., Song, G., Alström, P., & Fjeldså, J. (2015). The po‐ tential drivers in forming avian biodiversity hotspots in the east Himalaya Mountains of southwest China. Integrative Zoology, 10, 171–181. Li, S. H., Yeung, C. K., Feinstein, J., Han, L., Le, M. H., Wang, C. X., & Ding, P. (2009). Sailing through the Late Pleistocene: Unusual his‐ torical demography of an East Asian endemic, the Chinese Hwamei (Leucodioptron canorum canorum), during the last glacial period. Molecular Ecology, 18(4), 622–633. Librado, P., & Rozas, J. (2009). Dnasp v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics, 25, 1451–1457. Liu, K. B. (1988). Quaternary history of the temperate forests of China. Quaternary Science Review, 7, 1–20. Liu, S. X., Jiang, N., Xue, D. Y., Cheng, R., Qu, Y. H., Li, X. X., … Han, H. X. (2016). Evolutionary history of Apocheima cinerarius (Lepidoptera: Geometridae), a female flightless moth in northern China. Zoologica Scripta, 45(2), 160–174. Lou, M., & Golding, G. B. (2010). Assigning sequences to species in the absence of large interspecific differences. Molecular Phylogenetics and Evolution, 56, 187–194. Matsubayashi, K. W., Ohshima, I., & Nosil, P. (2010). Ecological speciation in phytophagous insects. Entomologia Experimentalis et Application, 134, 1–27. Meier, R., Shiyang, K., Vaidya, G., & Ng, P. K. (2006). DNA barcoding and taxonomy in diptera: A tale of high intraspecific variability and low identification success. Systematics Biology, 55, 715–728. Mende, M. B., & Hundsdoerfer, A. K. (2013). Mitochondrial lineage sort‐ ing in action‐historical biogeography of the Hyles euphorbiae com‐ plex (Sphingidae, Lepidoptera), Italy. BMC Evolutionary Biology, 13(1), 83–89. 7496 | Mikkola, K., & Ståhls, G. (2008). Morphological and molecular taxonomy of Dendrolimus sibiricus Chetverikov stat. rev. and allied lappet moths (Lepidoptera: Lasiocampidae), with description of a new species. Entomologica Fennica, 19, 65–85. Moritz, C., & Cicero, C. (2004). DNA barcoding: Promise and pitfalls. PLoS Biology, 2(10), e354. Peterson, M. A., & Denno, R. F. (1998). The influence of dispersal and diet breadth on patterns of genetic isolation by distance in phytophagous insects. The American Naturalist, 152(3), 428–446. Phillips, S. J., Anderson, R. P., & Schapire, R. E. (2006). Maximum entropy modeling of species geographical distributions. Ecological Modelling, 190, 231–259. Qiu, Y. X., Fu, C. X., & Comes, H. P. (2011). Plant molecular phylogeog‐ raphy in China and adjacent regions: Tracing the genetic imprints of Quaternary climate and environmental change in the world’s most diverse temperate flora. Molecular Phylogenetics and Evolution, 59(1), 225–244. Rach, J., DeSalle, R., Sarkar, I. N., Schierwater, B., & Hadrys, H. (2008). Character‐based DNA barcoding allows discrimination of genera, species and populations in Odonata. Proceedings of the Royal Society B: Biological Sciences, 275, 237–247. Rogers, A. R. (1995). Genetic evidence for a Pleistocene pop‐ ulation explosion. Evolution, 49, 608–615. https​://doi. org/10.1111/j.1558-5646.1995.tb022​97.x Rogers, A. R., & Harpending, H. (1992). Population growth makes waves in the distribution of pairwise genetic differences. Molecular Biology and Evolution, 9, 552–569. Ronquist, F., & Huelsenbeck, J. P. (2003). MRBAYES 3: Bayesian phyloge‐ netic inference under mixed models. Bioinformatics, 19, 1572–1574. Schneider, S., & Excoffier, L. (1999). Estimation of past demographic pa‐ rameters from the distribution of pairwise differences when the mu‐ tation rates vary among sites: Application to human mitochondrial DNA. Genetics, 152, 1079–1089. Singer, M. C., Ng, D., & Moore, R. A. (1991). Genetic variation in ovi‐ position preference between butterfly populations. Journal of Insect Behaviour, 4, 531–535. Tajima, F. (1989). Statistical method for testing the neutral mutation hy‐ pothesis by DNA polymorphism. Genetics, 123, 585–595. Tian, S., López‐Pujol, J., Wang, H., Ge, S., & Zhang, Z. Y. (2010). Molecular evidence for glacial expansion and interglacial retreat during Quaternary climatic changes in a montane temperate pine (Pinus kwangtungensis Chun ex Tsiang) in southern China. Plant Systematics and Evolution, 284, 219–229. Valade, R., Kenis, M., Hernandez‐Lopez, A., Augustin, S., Mena, N. M., Magnoux, E., … Lopez‐Vaamonde, C. (2009). Mitochondrial and mi‐ crosatellite DNA markers reveal a Balkan origin for the highly in‐ vasive horse‐chestnut leaf miner Cameraria ohridella (Lepidoptera, Gracillariidae). Molecular Ecology, 18, 3458–3470. Wang, J., Gao, P. X., Kang, M., Lowe, A. J., & Huang, H. W. (2009). Refugia within refugia: the case study of a canopy tree (Eurycorymbus cavaler‐ iei) in subtropical China. Journal of Biogeography, 36(11), 2156–2164. LI et al. Wang, G. D., Zhang, B. L., Zhou, W. W., Li, Y. X., Jin, J. Q., Shao, Y., … Che, J. (2018). Selection and environmental adaptation along a path to speciation in the Tibetan frog Nanorana parkeri. Proceedings of the National Academy of Sciences of the United States of America, 115(22), 5056–5065. Webb, I. I. I., & Bartlein, P. J. (1992). Global changes during the last 3 mil‐ lion years: Climatic controls and biotic responses. Annual Review of Ecology and Systematics, 23(1), 141–173. Yang, J. Q., Tang, J. Q., Liao, T. Y., Sun, Y., Zhou, Z. C., Han, C. C., … Lin, H. D. (2012). Phylogeographical analysis on Squalidus argentatus re‐ capitulates historical landscapes and drainage evolution on the is‐ land of Taiwan and mainland China. International Journal of Molecular Sciences, 13(2), 1405–1425. Ye, Z., Zhu, G. P., Chen, P., Zhang, D. L., & Bu, W. J. (2014). Molecular data and ecological niche modelling reveal the Pleistocene history of a semi‐aquatic bug (Microvelia douglasi douglasi) in East Asia. Molecular Ecology, 23, 3080–3096. Yu, G., Chen, X., Ni, J., Cheddadi, R., Guiot, J., Han, H., … Zheng, Z. (2000). Palaeovegetation of China: A pollen databased synthesis for the mid‐Holocene and last glacial maximum. Journal of Biogeography, 27, 635–664. Zhang, A. B., Li, D. M., Chen, J., & Zhang, Z. (2004). Geohistories of Dendrolimus punctatus (Walker) and its host plant pine genus in China. Chinese Journal of Applied Entomology, 41(2), 146–150. Zhang, A. B., Wang, Z. J., Tan, S. J., & Li, D. M. (2003). Monitoring the masson pine moth, Dendrolimus punctatus (Walker). (Lepidoptera: Lasiocampidae) with synthetic sex pheromone‐baited traps in Qianshan County, China. Applied Entomology and Zoology, 38(2), 177–186. Zhang, R., Song, G., Qu, Y., Alstrom, P., Ramos, R., Xing, X., … Lei, F. M. (2012). Comparative phylogeography of two widespread magpies: Importance of habitat preference and breeding behavior on ge‐ netic structure in China. Molecular Phylogenetics and Evolution, 65(4), 562–572. Zhao, Q. S., Wu, W. B., Lv, G. P., Chen, T. F., & Lin, Q. Y. (1999). Hybridization among pine moths species and its genetic rule. Scientia Silvae Sinicae, 35(4), 45–50. Zhou, Y. F., Abbott, R. J., Jiang, Z. Y., Du, F. K., Milne, R. I., & Liu, J. Q. (2010). Gene flow and species delimitation: A case study of two pine species with overlapping distributions in southeast China. Evolution, 64, 2342–2352. How to cite this article: Li J, Jin Q, Zhu G‐P, Jiang C, Zhang A‐B. Phylogeography of Dendrolimus punctatus (Lepidoptera: Lasiocampidae): Population differentiation and last glacial maximum survival. Ecol Evol. 2019;9:7480–7496. https​://doi. org/10.1002/ece3.5278

相关文章