Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Mitochondrial DNA markers reveal high genetic diversity and strong genetic differentiation in populations of Dendrolimus kikuchii Matsumura (Lepidoptera: Lasiocampidae)

  • Qiulei Men ,

    menqiulei888@126.com

    Affiliation School of Life Sciences, Provincial Key Laboratory of the Biodiversity Study and Ecology Conservation in Southwest Anhui, Research Center of Aquatic Organism Conservation and Water Ecosystem Restoration in Anhui Province, Anqing Normal University, Anqing, Anhui, P. R. China

  • Guoxi Xue,

    Affiliation School of Food and Bioengineering, Zhengzhou University of Light Industry, Zhengzhou, Henan, P. R. China

  • Dan Mu,

    Affiliation School of Life Sciences, Provincial Key Laboratory of the Biodiversity Study and Ecology Conservation in Southwest Anhui, Research Center of Aquatic Organism Conservation and Water Ecosystem Restoration in Anhui Province, Anqing Normal University, Anqing, Anhui, P. R. China

  • Qingling Hu,

    Affiliation School of Chemistry and Environment, Weinan Normal University, Weinan, Shaanxi, P. R. China

  • Minyi Huang

    Affiliation School of Life Sciences, Provincial Key Laboratory of the Biodiversity Study and Ecology Conservation in Southwest Anhui, Research Center of Aquatic Organism Conservation and Water Ecosystem Restoration in Anhui Province, Anqing Normal University, Anqing, Anhui, P. R. China

Abstract

Dendrolimus kikuchii Matsumura, 1927 is a serious forest pest causing great damage to coniferous trees in China. Despite its economic importance, the population genetics of this pest are poorly known. We used three mitochondrial genes (COI, COII and Cytb) to investigate the genetic diversity and genetic differentiation of 15 populations collected from the main distribution regions of D. kikuchii in China. Populations show high haplotype and nucleotide diversity. Haplotype network and phylogenetic analysis divides the populations into three major clades, the central and southeastern China (CC+SEC) clade, the eastern China (EC) clade, and the southwestern China (SWC) clade. Populations collected from adjacent localities share the same clade, which is consistent with the strong relationship of isolation by distance (r = 0.74824, P = 0.00001). AMOVA analysis indicated that the major portion of this molecular genetic variation is found among the three groups of CC+SEC, EC and SWC (61.26%). Of 105 pairwise FST comparisons, 93 show high genetic differentiation. Populations of Puer (PE), Yangshuo (YS) and Leishan (LS) are separated from other populations by a larger genetic distance. Distributions of pairwise differences obtained with single and combined gene data from the overall populations are multimodal, suggesting these populations had no prior population expansion in southern China. The nonsignificant neutral test on the basis of Tajima’ D and Fu’s Fs, and the lack of a star-shaped haplotype network together with the multiple haplotypes support this hypothesis. Pleistocene climatic fluctuations, combined with the host specificity to Pinus species, made these regions of south China into a refuge for D. kikuchii. The high level of population genetic structuring is related to their weak flight capacity, their variations of life history and the geographic distance among populations.

Introduction

Simao pine moth, Dendrolimus kikuchii Matsumura, 1927, is a destructive forest pest with an extensive range across southern China. Larvae attack various coniferous trees and regular outbreaks occur. These outbreaks are mainly on Pinus langhianensis Chev, P. yunnanensis Franch, P. massoniana Lambert, P. armandi Franch and Keteleeria evelyniana Mast. [1,2]. D. kikuchii larvae consume, on average, 7486.6 mm of P. langhianensis pine needles to complete their development [3]. During outbreaks, a large amount of pine needles are consumed, giving the appearance of forest fire damage. Larval damage from a large infestation of D. kikuchii can reduce the yield of timber, resin and cones, and affect the growth rate of pines, even resulting in tree death [1,2]. The Simao pine moth shows differences in ecological preferences among areas in China. One generation of D. kikuchii occurs in areas with short periods of optimal environment, such as in the middle region of Yunnan and Guizhou provinces. In Zhejiang, Fujian, and the southwest region of Yunnan provinces, two generations occur. In different counties of Yunnan province, such as Jingdong and Anning, the life histories of D. kikuchii also show significant variation [1].

Genetic diversity and genetic structure in insects can be affected by many factors, such as host plant species, chemical controls, geographic distance and geographic barriers [411]. For some lepidopterous species, the genetic diversity and genetic structure are related to their migration capacity and number of generations [1215]. Despite the economic and landscape threats of the D. kikuchii to pine trees in the southern parts of China and the need to establish control strategies for this pest, it has been unclear whether analogous associations between these factors and genotype patterns may be present among D. kikuchii populations.

Pleistocene climatic fluctuations are thought to have a great effect on this species’ distribution and historical demography [16]. South China was considered as a key area of some refugia for many relict rare species during the Pleistocene glaciation [17,18]. Zhang et al [19] studied the geohistory of Dendrolimus punctatus, a sympatric species of D. kikuchii, based on these geological events. They found that D. punctatus settled down in south China with the spread of masson pine during the Pleistocene. However, to our knowledge, there is no published report on the population history of D. kikuchii, and also no studies on the population history of Dendrolimus species based on molecular data.

In this study, we use three mitochondrial genes to (i) investigate the genetic diversity and genetic differentiation of 15 populations collected from the main distribution regions of D. kikuchii in southern China, and infer the demographic history of this pest, and (ii) to test the hypothesis that geographical isolation and biological characters, such as life history, are significant factors underlying genetic variation in Chinese D. kikuchii.

Material and methods

Ethics statement

There is no endangered or protected species involved in this study. No specific permissions were required for the described field studies for this widespread forest pest. We confirm that the locations are not privately owned or otherwise protected.

Sampling

A total of 182 individuals were collected from 15 locations during 2013 to 2016 within the D. kikuchii China distribution range (Table 1 and Fig 1). Different instar larvae and adult moths were sampled. For larvae, only one individual was collected per tree. For adult moths, pheromone traps were used in the pine forest (>1 ha) and only one moth from each trap was sampled in order to avoid sampling errors. All sampled insects were immersed in absolute ethyl alcohol, and then stored at -20°C prior to genetic analysis.

thumbnail
Fig 1. Sample locations for the 15 D. kikuchii populations from China, the codes for D. kikuchii populations are explained in Table 1; the gray region represents the geographic distribution of five D. kikuchii populations in Anhui.

https://doi.org/10.1371/journal.pone.0179706.g001

DNA extraction and amplification

Genomic DNA was extracted from the last proleg of the larvae or from any leg of the adult using the DNeasy Tissue Kit (QIAGEN, Hilden, Germany). Extraction was performed according to the bench protocol for animal tissues. Mitochondrial COI (LCO1490, HCO2198, [20]), COII (Eva, Patrick, [21]), and Cytb (generated for codling moth, unnamed [22]) were selected for use in this study. Polymerase chain reactions (PCR) were performed using a S1000 Thermal Cycler (BIO-RAD, Hercules, CA, USA) in a total volume of 20 μl, containing 10 μl 2×PCR Super Master Mix (Biotool, Shanghai, China), 0.25 μM of each primer, and 1 μl genomic DNA (10–30 ng/μl). PCR amplification was employed with denaturation at 95°C for 10 min, followed by 40 amplification cycles consisting of 95°C for 30 s, primer-specific annealing temperature of 53°C (COI), 52°C (COII and Cytb) for 1 min, 72°C for 45 s, and then a final step at 72°C for 10 min. Amplified products were purified and sequenced by Tianyi Huiyuan Biotechnology Co., Ltd.

Data analysis

The sequences were preliminarily aligned using the CLUSTAL X program [23]. Sequences of COI (646), COII (675), Cytb (700) of D. kikuchii were deposited in the NCBI GenBank (GenBank accession numbers: COI, MF155667-MF155697; COII, MF155698-MF155737; Cytb, MF155738-MF155760) (data in S1S3 Text). Multiple sequences of COI, COII and Cytb were concatenated to yield a total length of 2021 bp. The haplotype network of D. kikuchii was analyzed using a median-joining algorithm in the program Network 4.6 [24]. A neighbour-joining (NJ) tree was built using NJ tree subroutine in NEIGHBOUR within PHYLIP 3.5 [25], and the parameters were expanded 1,000 times. The CONSENSE subroutine within PHYLIP was then applied to generate a consensus NJ tree that provided bootstrap support at each node. The tree was visualized using treeview version 1.6.6 software. DnaSP 5.0 [26] were performed to calculate number of polymorphic sites (S), number of haplotypes (H), haplotype diversity (Hd), nucleotide diversity (Pi), Tajima’s D (D), and Fu’s Fs (Fs). Analysis of molecular variance (AMOVA) was performed using the ARLEQUIN version 3.5 based on the combinations of the three gene sequences [27], along with calculating pair fixation indices (FST). The pairwise genetic distances were calculated by MEGA 6.0 [28] based on the Kimura-2-parameter model [29]. Referring to the criterion for genetic differentiation by Wright (1978) [30], we defined genetic differentiation as low for FST<0.05, moderate for 0.05<FST<0.15, high for 0.15<FST<0.25, and very high for FST>0.25 [31]. In order to test isolation by distance (IBD), the matrices of genetic distance FST/(1-FST) and the geographic distance (ln) between all 15 sampling populations were compared using the Mantel test with 10,000 permutations [32]. This analysis was performed using the ZT software package [33]. For examining demographic history, the distribution of pairwise differences between individual sequences was analyzed by means of mismatch distribution analysis using DnaSP 5.0 [26]. The formula, Tau = 2 ut was used to detect the time of population size changes [34]. The nucleotide substitution rate in mitochondrial DNA was 2.3% per million years (MY) as suggested in Knowles et al. (2000) [35].

Results

Genetic diversity

For concatenated sequences, the haplotype diversity ranged from 0 to 1 with a average of 0.940, while the nucleotide diversity ranged from 0 to 0.00251 with a average of 0.01451 (Table 2). All populations displayed large numbers of mitochondrial haplotypes, with a total 31 haplotypes obtained for COI, 40 haplotypes for COII, and 23 haplotypes for Cytb. Seven, six and four common haplotypes were shared for COI, Cytb and COII respectively (Fig 2). The median-joining network demonstrated a high genetic diversity for populations of D. kikuchii in China. Of 73 examined haplotypes, 68 were unique. Most populations lacked a common haplotype.

thumbnail
Fig 2. Median-Joining network based on the single genes of COI, COII and Cytb haplotypes.

Each circle represents a haplotype, and the area of a circle is proportional to the number of observed individuals. Colors within the nodes refer to the D. kikuchii sampling regions.

https://doi.org/10.1371/journal.pone.0179706.g002

thumbnail
Table 2. Parameters of genetic diversity based on mitochondrial sequence data of 15 populations of D. kikuchii.

https://doi.org/10.1371/journal.pone.0179706.t002

Genetic differentiation

The Median-Joining network of the haplotypes can be divided into three major clades (Clades CC+SEC, EC and SWC) (Fig 3). All haplotypes in Clade CC+SEC were obtained from samples from central China and southeastern China including Hunan, Jiangxi, and Fujian provinces, while all haplotypes in Clade EC were from eastern China including Anhui and Zhejiang provinces. The haplotypes in Clade SWC included samples from southwestern China and covered Yunnan, Guizhou and Guangxi provinces. Overall, populations collected from adjacent localities or the same province shared the same clade. However, the Median-Joining network of the haplotypes based on single gene of COI and Cytb was not in line with the result from the combined genes. This may due to less nucleotide variation in both single genes.

thumbnail
Fig 3. Median-Joining network based on the combined gene of COI, COII and Cytb haplotypes.

Each circle represents a haplotype, and the area of a circle is proportional to the number of observed individuals. Colors within the nodes refer to the D. kikuchii sampling regions. A, B, and C indicate the three clades obtained.

https://doi.org/10.1371/journal.pone.0179706.g003

Phylogenetic reconstruction using the Kimura-2-parameter resulted in a consensus NJ tree with comparatively higher bootstrap values. The populations were divided into three major clusters, CC+SEC, EC and SWC, which comply well with the results from the haplotype network (Fig 4).

thumbnail
Fig 4. Unrooted NJ phylogenetic tree of D. kikuchii based on the combination of the three gene sequences, estimated with PHYLIP using Kimura-2-parameter (1000 replications).

The bootstraps were shown near the nodes. A, southwestern China clade; B, central and southeastern China clade; C, eastern China clade.

https://doi.org/10.1371/journal.pone.0179706.g004

The values of pairwise FST range from 0 to 0.98921. Of 105 comparisons, 93 showed high genetic differentiation. The pairwise FST values between JA and JJ populations were less than 0.15, indicating moderate genetic differentiation. The pairwise FST values between the populations from Anhui and Zhejiang provinces, ranged from 0 to 0.06363, indicating low genetic differentiation. FST values among populations from Anhui Province range from 0 to 0.11845, suggesting a relatively low genetic differentiation (Table 3). This is consistent with the results of the clustering analysis based on a Median-Joining network.

thumbnail
Table 3. Pairwise FST (below diagonal) and genetic distance (above diagonal) based on mitochondrial sequence data of 15 populations of D. kikuchii.

https://doi.org/10.1371/journal.pone.0179706.t003

The values of pairwise genetic distance between PE population and other populations range from 0.032 to 0.037, indicating high genetic differentiation.

Similar high values were observed in comparisons of the LS population with other populations, which range from 0.016 to 0.032. Comparisons of the YS population with other populations range from 0.015 to 0.032. The values of genetic distance among the remaining populations are less than 0.01 (Table 3).

AMOVA results indicate that the major portion of the molecular genetic variation is found among groups (61.26%). Exact tests showed a significant genetic variance on all three levels (P<0.001) (Table 4).

The Mantel test for the 15 populations revealed a positive correlation between genetic distances and geographic distances (r = 0.74824, P = 0.00001), suggesting that isolation by distance had a limiting effect on gene flow.

Mismatch distribution

The results of the combined gene analysis show that Tajima’ s D values are significantly positive with a value of 0.24250, but are not significant in most specific populations (P<0.05 in HS, PE and HY, P>0.05 in the rest of the populations). Fu’s F statistic was significantly negative with a value of -5.132 (P>0.1) (Table 5). At the population level, HS, HY and PE populations have negative and significant Tajima’ s D and Fu’s Fs values, whereas HS, HY and PE populations show a bimodal distributions, suggesting that expansion events were not detected with this analysis (data not shown). Distributions of pairwise differences (mismatch distributions) obtained with the single and combined gene data from the overall populations were multimodal, suggesting that the populations of D. kikuchii in southern China did not experience population expansion (Fig 5). A nonsignificant neutral test based on Tajima’ D and Fu’s Fs support this interpretation. The time of reaching a stable population size are estimated to be 26500 (COI), 21600 (COII), and 11800 (Cytb) years ago.

thumbnail
Fig 5. Observed and expected mismatch distributions for D. kikuchii in China, based on COI, COII, Cytb, and combined gene sequences.

https://doi.org/10.1371/journal.pone.0179706.g005

thumbnail
Table 5. Parameters of the neutral test based on mitochondrial sequence data of 15 populations of D. kikuchii.

https://doi.org/10.1371/journal.pone.0179706.t005

Discussion

Using three mitochondrial genes, we investigated the genetic diversity and structure of 182 individuals of 15 D. kikuchii populations sampled throughout their main areas of distribution in China. The results show a high genetic diversity and high level of genetic structuring of D. kikuchii in the sampled areas.

All populations displayed large numbers of mitochondrial haplotypes, with a total 31 haplotypes obtained for COI, 40 haplotypes for COII, and 23 haplotypes for Cytb, of which seven, six and four were common haplotypes shared respectively. Of 73 examined haplotypes based on combined genes, 68 were unique and did not share the same ancestral haplotype. High numbers of private haplotypes and lack of ancestral haplotype suggest that D. kikuchii could be a species native to China. In a previous study, Zheng et al. [36] indicated that these two features can be considered as indicators for native species, especially for native species with a low dispersal capacity, such as Grapholita molesta Busck (Lepidoptera: Tortricidae) and Chilo suppressalis (Walker) (Lepidoptera: Pyralidae). Although the origin of D. kikuchii remains uncertain, the higher genetic diversity we observed in the present study supports our assumption that China is part of the original range of D. kikuchii.

Based on the results of three mitochondrial genes, including AMOVA analysis, haplotype network and phylogenetic analysis, we conclude that D. kikuchii populations were geographically structured in three regions: eastern China, southwestern China, and central China along with southeastern China. The sampling localities of southwestern China are isolated by the Wumeng, Leigong, and Fanjingshan Mountains, while sampling localities of central China are isolated by the Hengshan and Dabieshan Mountains, which have acted as substantial barriers to gene flow. Populations from central China and southeastern China shared common haplotypes; this is in accordance with the results of phylogenetic analysis. Although there is a geographical barrier formed by Wuyishan Mountain between these two regions, a transportation network increases the gene flow among D. kikuchii populations. In Lepidoptera species, dispersal patterns also influence genetic variation [37,38]. D. kikuchii is generally regarded as a sedentary species based upon previous studies of its flight capacity, with a sphere of activity extending over 10~20 km [1, 3]. The weak flight capacity of D. kikuchii can reduce gene flow among populations. The IBD relationship (r = 0.74824, P = 0.00001) in the present study supports this hypothesis. Similar trends have been identified for many sedentary species, such as Chilo suppressalis (Walker) [15] and Carposina sasakii Matsumura [18]. The same result was reported by Weng et al. [39]. They found that the genetic variation of D. punctatus populations that ranged over five adjacent regions in Zhejiang province were low using the ISSR-PCR marker.

Analysis of genetic distance indicate that populations of Puer (PE), Yangshuo (YS) and Leishan (LS) are separated from other populations. In fact, D. kikuchii caterpillars in Puer and Yangshuo populations turn into adults one moth earlier than those of other populations. And in Guizhou, only one generation occurs, while two generations occur in the rest of the provinces [1]. These variations in life history may contribute to the significant genetic differentiation. Moreover, the areas of distribution of D. kikuchii in China occur across very complex topography (tall mountains, plain and basin), different climates (temperate and tropical climates), different agricultural landscapes and forest types. A high level of population differentiation presents a high potential for adaptation to different environmental conditions [40] and high reproductive rates [41], which allow the moth to form locally differentiated populations. Similar results have been found for Chilo suppressalis [15] and Reticulitermes chinensis [42].

Distributions of pairwise differences (mismatch distributions) obtained with COI, COII, Cytb, and combined gene data from the overall populations are multimodal, suggesting that the populations of D. kikuchii did not experience population expansion. The lack of a star shape for the haplotype network together with the existence of multiple haplotypes support such a hypothesis. This also explains why there are so many private haplotypes in different geographical populations. The times when population histories stabilized are estimated to be 26500 (COI), 21600 (COII), and 11800 (Cytb) year ago. This range is within the late Pleistocene, which was characterized by climatic oscillations between warm and cold periods [43]. Over 42 thousand years ago, Pinus massoniana, the favorite host plant for D. kikuchii, gradually spread into the warmer south China, a condition they would prefer [19]. These events impacted the distribution of D. kikuchii, with these regions of south China gradually becoming the refuge for D. kikuchii. During the late Pleistocene, some regions of south China were not covered by large ice sheets [44]. In addition, south China comprises a mountainous mosaic area and has the potential to host microclimatic zones that are probably capable of providing some key refuges for many relict species [45]. Considering the lack of an ancestral haplotype and strong isolation-by-distance relationships of this species, we can conclude that D. kikuchii in south China has arisen in separate refuges and experienced parallel evolutions. Similar results have been found for grasshoppers [46], and the blue manakin [47].

Future population genetic research on D. kikuchii in China should cover a larger area and larger number of sampled individuals, as well as use nuclear genes as markers to provide a greater understanding of genetic structure. Additional research needs to be done to detail the geographic origins of D. kikuchii in China and its spread through China.

Conclusion

Using three mitochondrial genes, we investigated the genetic diversity and structure of 182 individuals of D. kikuchii sampled throughout its main distribution areas in China. The results show high genetic diversity and a high level of genetic structuring of D. kikuchii in sampled areas. The high level of population genetic structuring is related to the weak flight capacity of the D. kikuchii, variations in its life history and the geographic distance among populations. Distributions of pairwise differences (mismatch distributions) obtained with COI, COII, Cytb, and combined gene data indicate the populations of D. kikuchii in southern China did not experience population expansion. These genetic data not only provide us with an understanding of population genetics for such a secondary species, but also provide guidance for pest management strategies.

Supporting information

S1 Text. Submission data of COI gene segment (numbered 2018976) in NCBI GenBank.

https://doi.org/10.1371/journal.pone.0179706.s001

(TXT)

S2 Text. Submission data of COII gene segment (numbered 2018977) in NCBI GenBank.

https://doi.org/10.1371/journal.pone.0179706.s002

(TXT)

S3 Text. Submission data of Cytb gene segment (numbered 2018978) in NCBI GenBank.

https://doi.org/10.1371/journal.pone.0179706.s003

(TXT)

Acknowledgments

We thank Dr. John Richard Schrock, Emporia State University, Emporia, Kansas, U.S.A., for correcting the language in an early draft of the manuscript. We thank Qing Tong and Yue Liu (Puer Forestry Research Institute, Puer, Yunnan), Delan Chen (Wuyishan Forestry Bureau, Wuyishan, Fujian), Xiangbo Kong (Chinese Academy of Forestry Sciences, Beijing), Shunli Chen (Fujian Agriculture and Forestry University, Fuzhou, Fujian), and Hongbin Zeng (Jian Forestry Bureau, Jian, Jiangxi) for providing insect samples.

Author Contributions

  1. Conceptualization: QLM GXX.
  2. Data curation: QLM GXX.
  3. Formal analysis: QLM GXX.
  4. Investigation: QLM GXX.
  5. Project administration: QLM DM MYH.
  6. Resources: QLM DM MYH.
  7. Writing – original draft: QLM GXX QLH.

References

  1. 1. Hou TQ. The Pine Caterpillars in China. Beijing, China: Science Press; 1987.
  2. 2. Dai QY, Gao Q, Wu CS, Chesters D, Zhu CD, Zhang AB. Phylogenetic reconstruction and DNA barcoding for closely related pine moth species (Dendrolimus) in China with multiple gene markers. PLoS ONE 2012; 7(4): e32544. pmid:22509245
  3. 3. Tong Q, He JZ. Study of biological characters of Dendrolimus kikuchii Matsumura and food consumption of its larva. Journal of Anhui Agricultural Science 2009; 37: 13122–13124.
  4. 4. Wei SJ, Shi BC, Gong YJ, Jin GH, Chen XX, Meng XF. Genetic structure and demographic history reveal migration of the diamondback moth Plutella xylostella (Lepidoptera: Plutellidae) from the southern to northern regions of China. PLoS ONE 2013; 8(4): e59654. pmid:23565158
  5. 5. Nason JD, Heard SB, Williams FR. Host-associated genetic differentiation in the golden rod elliptical-gall moth, Gnorimoschema gallaesolidaginis (Lepidoptera: Gelechiidae). Evolution 2002; 56 (7): 1475–1488. pmid:12206247
  6. 6. Franck P, Reyes M, Olivares J, Sauphanor B. Genetic architecture in codling moth populations: comparison between microsatellite and insecticide resistance markers. Molecular Ecology 2007; 16: 3554–3564. pmid:17845430
  7. 7. Du YL, Zhang JX, Yan ZG, Ma YQ, Yang MM, Zhang MZ, Zhang ZY, Qin L, Cao QQ. Host preference and performance of the yellow peach moth (Conogethes punctiferalis) on chestnut cultivars. PLoS ONE 2016; 11 (6): e0157609. pmid:27326856
  8. 8. Timm AE, Geertsema H, Warnich L. Gene flow among Cydia pomonella (Lepidoptera: Tortricidae) geographic and host populations in South Africa. Journal of Economic Entomology 2006; 99 (2): 341–348. pmid:16686131
  9. 9. Calcagno V, Thomas Y, Bourguet D. Sympatric host races of the European corn borer: adaptation to host plants and hybrid performance. Journal of Evolutionary Biology 2007; 20: 1720–1729. pmid:17714289
  10. 10. Miller NJ, Birley AJ, Overall ADJ, Tatchell GM. Population genetic structure of the lettuce root aphid, Pemphigus bursarius (L.), in relation to geographic distance, gene flow and host plant usage. Heredity 2003; 91: 217–223. pmid:12939621
  11. 11. Bouvier JC, Buès R, Boivin T, Boudinhon L, Beslay D, Sauphanor B. Deltamethrin resistance in the codling moth (Lepidoptera: Tortricidae): inheritance and number of genes involved. Heredity 2001; 82: 456–462.
  12. 12. Chen MH, Dorn S. Microsatellites reveal genetic differentiation among populations in an insect species with high genetic variability in dispersal, the codling moth, Cydia pomonella (L.) (Lepidoptera: Tortricidae). Bulletin of Entomological Research 2009; 100: 75–85. pmid:19366473
  13. 13. Men QL, Chen MH, Zhang YL, Feng JN. Genetic structure and diversity of a newly invasive species, the codling moth, Cydia pomonella (L.) (Lepidoptera: Tortricidae) in China. Biological Invasions 2013; 15: 447–458.
  14. 14. Bohonak AJ. Dispersal, gene flow and population structure. Quarterly Review of Biology 1999; 74: 21–45. pmid:10081813
  15. 15. Meng XF, Shi M, Chen XX. Population genetic structure of Chilo suppressalis (Walker) (Lepidoptera: Crambidae): strong subdivision in China inferred from microsatellite markers and mtDNA gene sequences. Molecular Ecology 2008; 17: 2880–2897. pmid:18482260
  16. 16. Fan Z, Jiang GF, Liu YX, He QX, Blanchard B. Population explosion in the Yellow-Spined bamboo locust Ceracris kiangsu and inferences for the impact of human activity. PLoS ONE 2014; 9(3): e89873. pmid:24603526
  17. 17. Zhan XJ, Zheng YF, Wei FW, Bruford WM, Jia CX. Molecular evidence for Pleistocene refugia at the eastern edge of the Tibetan Plateau. Molecular Ecology 2011; 20: 3014–3026. pmid:21689184
  18. 18. Wang J, Yu Y, Li LL, Guo D, Tao YL, Chu D. 2015. Carposina sasakii (Lepidoptera: Carposinidae) in its native range consists of two sympatric cryptic lineages as revealed by mitochondrial COI gene sequences. Journal of Insect Science 2015; 15(1): 85. pmid:26136498
  19. 19. Zhang AB, Li DM, Chen J, Zhang Z. The preliminary study on geohistory of Dendrolimus punctatus and its host plants, Pinus spp. Entomological Knowledge 2004; 41: 146–149.
  20. 20. Folmer O, Black M, Hoeh W, Lutz R, Vrijenhoek R. DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Molecular Marine Biology and Biotechnology 1994; 3: 294–299. pmid:7881515
  21. 21. Caterino MS, Sperling FAH. Papilio phylogeny based on mitochondrial cytochrome oxidase I and II genes. Molecular Phylogenetics and Evolution 1999; 11 (1): 122–137. pmid:10082616
  22. 22. Li YT, Duan XL, Qiao XF, Li XY, Wang K, Men QL, Chen MH. Mitochondrial DNA revealed the extent of genetic diversity and invasion origin of populations from two separate invaded areas of a newly invasive pest, Cydia pomonella (L.) (Lepidoptera: Tortricidae) in China. Bulletin of Entomological Research 2015; 105 (4): 485–496. pmid:25895900
  23. 23. Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG. The clustal_x Windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Research 1997; 24: 4876–4882.
  24. 24. Bandelt HJ, Forster P, Röhl A. Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution 1999; 16: 37–48. pmid:10331250
  25. 25. Felsenstein J. PHYLIP—phylogeny inference package (version 3.2). Cladistics 1989; 5: 164–166.
  26. 26. Bonnet E, Van der Peer Y. ZT: a software tool for simple and partial Mantel tests. Journal of Statistical Software 2002; 7: 1–12
  27. 27. Excoffier L, Lischer HL. Arlequin suite ver 3.5: a new series of programs to perform population genetics analyses under Linux and Windows. Molecular Ecology 2010; 10: 564–567.
  28. 28. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Molecular Biology and Evolution 2013; 30: 2725–2729. pmid:24132122
  29. 29. Kimura M. A simple method for estimating evolutionary rate of base substitutions through comparative studies of nucleotide sequences. Journal of Molecular Evolution 1980; 16: 111–120. pmid:7463489
  30. 30. Wright S. Evolution and Genetics of Populations. Chicago: University of Chicago; 1978.
  31. 31. Govindajuru RD. Variation in gene flow levels among predominantly self-pollinated plants. Journal of Evolutionary Biology 1989; 2: 173–181.
  32. 32. Mantel N. The detection of disease clustering and a generalized regression approach. Cancer Research 1967; 27: 209–220 pmid:6018555
  33. 33. Librado P, Rozas J. DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009; 25: 1451–1452. pmid:19346325
  34. 34. Rogers AR, Harpending H. Population growth makes waves in the distribution of pairwise genetic differences. Molecular Biology and Evolution 1992; 9: 552–569. pmid:1316531
  35. 35. Knowles LL. Test of Pleistocene speciation in montane grasshoppers (Genus Melanoplus) from the sky islands of western North America. Evolution 2000; 54: 1337–1348. pmid:11005300
  36. 36. Zheng Y, Peng X, Liu GM, Pan HY, Dorn S, Chen MH. High genetic diversity and structured populations of the oriental fruit moth in its range of origin. PLoS ONE 2013; 8 (11): e78476. pmid:24265692
  37. 37. Boivin T, Bouvier JC, Beslay D, Sauphanor B. Variability in diapause propensity within populations of a temperate insect species: interactions between insecticide resistance genes and photoperiodism. Biological Journal of Linnean Society 2004; 83: 341–351
  38. 38. Fuentes-Contreras E, Espinoza JL, Lavandero B, Ramírez CC. Population genetic structure of codling moth (Lepidoptera: Tortricidae) from apple orchards in Central Chile. Journal of Economic Entomology 2008; 101: 190–198. pmid:18330135
  39. 39. Weng HB, Niu BL, He LH, Jiang P, Meng ZQ. Genetic diversity among Dendrolimus punctatus populations sampled from different regions of Zhejiang Province. Scientia Silvae Sinicae 2008; 44 (1): 164–168.
  40. 40. Cardé RT, Minks AK. Control of moth pests by mating disruption: successes and constraints. Annual Review of Entomology 1995; 40: 559–585.
  41. 41. Myers CT, Hull LA, Krawczyk G. Comparative survival rates of oriental fruit moth (Lepidoptera: Tortricidae) larvae on shoots and fruit of apple and peach. Journal of Economic Entomology 2006; 99: 1299–1309. pmid:16937685
  42. 42. Huang QY, Li GH, Husseneder C, Lei CL. Genetic analysis of population structure and reproductive mode of the termite Reticulitermes chinensis Snyder. PLoS ONE 2013; 8(7): e69070. pmid:23894408
  43. 43. Knowles LL, Carstens BC, Keat ML. Coupled genetic and ecological niche-models to examine how past population distributions contribute to divergence. Current Biology 2007; 17: 1–7.
  44. 44. Hewitt GM. The genetic legacy of the Quaternary ice ages. Nature 2000; 405: 907–913. pmid:10879524
  45. 45. Shen L, Chen XY, Li YY. Glacial refugia and postglacial recolonisation patterns of organisms. Acta Ecology Sinica 2002; 22: 1985–1990.
  46. 46. Rosetti N, Remis MI. Spatial genetic structure and mitochondrial DNA phylogeography of Argentinean populations of the grasshopper Dichroplus elongatus. PLoS ONE 2012; 7(7): e40807. pmid:22859953
  47. 47. Francisco MR, Gibbs HL, Galetti M, Lunardi VO, Galettijunior PM. Genetic structure in a tropical lek-breeding bird, the blue manakin (Chiroxiphia caudata) in the Brazilian Atlantic Forest. Molecular Ecology 2007; 16: 4908–4918. pmid:17927710