Next Article in Journal
Climatic Variation of Supercooling Point in the Linden Bug Pyrrhocoris apterus (Heteroptera: Pyrrhocoridae)
Next Article in Special Issue
The Past, Present, and Future of the Hemlock Woolly Adelgid (Adelges tsugae) and Its Ecological Interactions with Eastern Hemlock (Tsuga canadensis) Forests
Previous Article in Journal
The Factors Influencing Seasonal Dynamics and Spatial Distribution of Stable Fly Stomoxys calcitrans (Diptera, Muscidae) within Stables
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Limited Genetic Structure of Gypsy Moth Populations Reflecting a Recent History in Europe

1
Croatian Forest Research Institute, Cvjetno naselje 41, 10450 Jastrebarsko, Croatia
2
UMR CNRS-UFC 6249 Chrono-Environment, Université de Franche Comte, 25200 Montbéliard, France
3
Faculty of Science, University of Zagreb, HR-10000 Zagreb, Croatia
4
Department of Forest and Soil Sciences, BOKU, University of Natural Resources and Life Sciences, A-1180 Vienna, Austria
5
Forest Research Institute, Hellenic Agricultural Organization Demeter, Vasilika, 57006 Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Insects 2018, 9(4), 143; https://doi.org/10.3390/insects9040143
Submission received: 13 August 2018 / Revised: 14 October 2018 / Accepted: 17 October 2018 / Published: 18 October 2018
(This article belongs to the Special Issue Ecology and Management of Forest Insects in a Changing World)

Abstract

:
The gypsy moth, Lymantria dispar, a prominent polyphagous species native to Eurasia, causes severe impacts in deciduous forests during irregular periodical outbreaks. This study aimed to describe the genetic structure and diversity among European gypsy moth populations. Analysis of about 500 individuals using a partial region of the mitochondrial COI gene, L. dispar was characterized by low genetic diversity, limited population structure, and strong evidence that all extant haplogroups arose via a single Holocene population expansion event. Overall 60 haplotypes connected to a single parsimony network were detected and genetic diversity was highest for the coastal populations Croatia, Italy, and France, while lowest in continental populations. Phylogenetic reconstruction resulted in three groups that were geographically located in Central Europe, Dinaric Alps, and the Balkan Peninsula. In addition to recent events, the genetic structure reflects strong gene flow and the ability of gypsy moth to feed on about 400 deciduous and conifer species. Distinct genetic groups were detected in populations from Georgia. This remote population exhibited haplotypes intermediate to the European L. dispar dispar, Asian L. dispar asiatica, and L. dispar japonica clusters, highlighting this area as a possible hybridization zone of this species for future studies applying genomic approaches.

1. Introduction

The gypsy moth, Lymantria dispar L. (Lepidoptera, Erebidae), is a polyphagous pest feeding on more than 400 plant species with a preference for species within the genus Quercus [1]. Lymantria dispar is native to the Palearctic region [1,2] and became invasive in North America after its intentional introduction for hybridization experiments [3]. Currently, L. dispar is considered one of the most notorious forest pests worldwide [4], with outbreaks in its native range being common particularly in the oak forests of the Mediterranean region, while in Asia the broader spectrum of potential hosts renders gypsy moth capable of even more massive outbreaks [5]. Nonetheless, the potential of gypsy moth populations to erupt varies greatly [6] as in cases of recent invasion an outbreak can be sustained [7]. For all these reasons, gypsy moth, has been in the spotlight of various investigations ranging from population dynamics [8] and dispersal ecology [9,10] to the distribution of its genetic diversity at a global scale [11,12]. Three subspecies are recognized throughout the temperate part of the world due to their economic importance; European (L. dispar dispar) and two Asian (L. dispar asiatica in Russia and Asia and L. dispar japonica in Japan), and each of them is described separately based on morphological and behavioral traits [1]. Nonetheless, the status of the two Asian subspecies still remains disputable [13], as recent molecular investigations argue for an even more elaborate pattern among the far east populations [12]. However, female flight capability turned out to be the most important behavioral trait distinguishing Asian subspecies from European [10,14].
First genetic analysis based on DNA sequences resulted in developing an mtDNA marker that verified the close resemblance of North American and European population, that both were distinctly separated from Asian populations [11]. This pattern was confirmed in subsequent studies which showed that North American populations were related to French populations and additionally that Asian population were separated into a western and eastern genetic entity [14], where Japanese populations form a separate cluster differentiated from Western Asian populations [15].
Until recently, studies on European populations of gypsy moth were mostly integrated into broader investigations that aimed at a global overview of this species [14,15]. A study analyzing Croatian populations revealed a genetic split between coastal and continental populations caused by the Dinaric Alps [16]. In this study we sampled 38 locations from sixteen European countries, and additionally two Georgian populations to reveal insight into the phylogeography and demography of this species in Europe.

2. Materials and Methods

Samples (larvae) were collected from 36 localities in sixteen European countries and two additional sites from Georgia between 2010 and 2014 (Table 1), with specimens from the same locality/site constituting a geographic population. Larvae were taken from distant trees (not more than one larva per tree) so to avoid any bias induced by the use of mtDNA marker, and after collection they were immediately put into 99% ethanol and stored at −20 °C in laboratory. In total, 497 individuals were analyzed. One pair of abdominal legs was used for genomic DNA extraction from each specimen individually, using SIGMA Aldrich mammalian genomic DNA extraction kit (Sigma, St. Louis, MO, USA) and following manufacturer’s instruction. DNA amplification was performed with primers UEA5 and UEA10 [17] in 20 μL reactions containing 2 mM of MgCl2, 100 mM of dNTPs (Fermentas, Vilnius, Lithuania), 0.5 mM of each primer 1U of Taq (Fermentas, Vilnius Lithuania). Thermocycling consisted of an initial denaturation step at 94 °C (2 min), which was followed by 33 cycles at 94 °C (30 s), 48 °C (60 s), and 68 °C (60 s), and a final extension step at 68 °C (10 min). All reactions were checked for amplification by gel electrophoresis, and confirmed products were purified using peqGOLD Cycle-Pure Kit (PeqLab, Erlangen, Germany) and directly sequenced with UEA10 at Eurofins MWG Operon (Mainz, Germany). Sequences were manually examined using Chromas Lite (Technelysium, Brisbane Australia) and aligned using ClustalW algorithm incorporated in BioEdit [18]. Upon alignment, ends of sequences were trimmed in order to obtain full overlap. Finally, this alignment was used for determination of haplotype sequences, using TCS 1.21 [19]. Singleton sequences were verified by an additional PCR and sequencing. Haplotype sequences were deposited in NCBI with accession numbers: KY628707 to KY628749.
Haplotype diversity, nucleotide diversity, mean number of pairwise differences and allelic richness were calculated for each population using ARLEQUIN [20] and Contrib 1.02 [21]. Parsimony network was reconstructed by cladistic analysis in TCS 1.21 and drawn in CorelDraw X5 (Ottawa, ON, Canada). Ambiguities in the cladogram were resolved by applying topological, geographic, and frequency criteria, where applicable [22], while phylogeographic inferences were assessed with ANeCA v1.2 [23]. Overall population differentiation was assessed by testing if NST(obs) was greater than NST(exp) using heuristic permutation test with 1000 replicates in PERMUT 1.0 [24], and subsequently compared with GST in order to assess the depth of differentiation.
To separate the source of variation in three hierarchical levels (among groups of populations (FCT), among populations within groups (FST), and within populations (FSC)), objective grouping and Analysis of Molecular Variance were performed using ARLEQUIN [19] based on the host of each population (Table 1). The best-fit nucleotide substitution model of our data was selected among 88 different models (in this case GTR with Gamma distributed heterogeneity of rates) as implemented in jModeltest 2.1.7 [25], and this model was then used in MrBayes 3.2 [26]. Two separate chains were run in the Bayesian analysis, and the initial run lasted 2 × 106 generations, with a sampling frequency of 100 generations in order to define the trees to be discarded in the second run. As a plateau between standard deviations was reached after 5 × 105 generations, 500 trees were discarded from the estimation of the consensus tree. To better understand the assignment of haplotypes in clades and their geographic distribution, we also used sequences deposited in NCBI GenBank from previous investigations of L. dispar (HM013724 (Russia, Vostochnyy), HM013736 (Japan), HM013737 (Japan) [15]; KX436205 (Japan, Honshu), KX436223 (Japan, Nagano), KX436230 (Russia, Vladivostok) [27]; KY923059 (Russia, Primorski), KY923063 (Lithuania), KY923064 (Russia, Krasnoyarsk), KY923065 (Kazhakstan), GU994784 (Mongolia) [28]).
Locality map and haplotype distribution map were constructed using ArcMap 9. 3 (ESRI, Redlands, California, USA) and Genetic Landscape Toolbox [29] and arranged in CorelDraw X5 (Ottawa, ON, Canada). Finally, haplotype rarefaction analysis was constructed using R3.2.3 [30] with the function specaccum of the vegan package [31].

3. Results

In total, 497 sequences 676 bp long (spanning from 699 to 1375 bp of L. dispar COI sequence) were used for further analyses. Cladistic analysis resulted in 60 haplotypes (Figure 1) distinguished by 25 singletons and 19 parsimony informative polymorphic sites, with the rarefaction curve of haplotypes reaching a plateau (data not shown) showing that the sampling effort was sufficient to describe intraspecific haplotype diversity [32]. Total haplotype diversity HT was 0.81 and intrapopulation diversity HS was 0.437 (data not shown). The lowest genetic diversity values were spotted in populations from the continental basin of the Balkan Peninsula between Dinaric Alps, Alps and Carpathian Alps and in northern Continental regions. On the other hand, the highest genetic diversity was observed in Coastal Croatia, Italy, and France. Populations from the Aegean coast and Black sea coast showed rather low diversity, while populations from Georgia showed intermediate values (Table 1).
Fixation index GST (se) = 0.463 (0.056) was smaller than observed fixation index NST(obs.) (se) = 0.521 (0.059) which was significantly greater than theoretical fixation index NST(theor.) (se) = 0.456 (0.001) (p = 0.010), confirming the existence of a limited phylogeographic structure (GST < NST(obs.) > NST(theor.)) (Figure 2). Nevertheless, AMOVA attributed not more than 24% of divergence to host selection (FCT=0.23831, p < 0.01), with the main source of divergence being found among populations within groups (FST = 0.53239, p < 0.01), and within populations (FSC = 0.38609, p < 0.01), rejecting the hypothesis of host-driven divergence at least among the broadleaved hosts analyzed in the current study.
The phylogenetic tree grouped European haplotypes in a similar pattern supporting the main clades with posterior probabilities higher than 80%, with the exception of green cluster that was the most common one in Central Europe. The inclusion of NCBI GenBank sequences from Asia, revealed the position of haplotypes LD52, LD53, and LD54 from Georgia diverged 1.46% from the European groups, while averagely diverged 0.86% from the cluster of Asian haplogroups L. dispar asiatica/L. japonica haplotypes (KX436223, KY923509, HM013724, GU994784, HM013737, HM013736, KX436205, KX436230) retrieved from NCBI (Figure 3). Finally, while Siberian sequences integrated in the European L. dispar dispar cluster, Georgian haplotypes formed a distinct clade, separated from L. dispar japonica and L. dispar asiatica.

4. Discussion

Gypsy moth is a holarctic species, flexible in terms of temperature tolerance [34] and it has extremely wide spectrum of host plants [35]. This species successfully invaded a vast variety of ecosystems, causing disturbances on a wide array of tree and general plant species. Model predictions for L. dispar dispar support a range shift of up to 900 km northwards in Europe as a consequence of increase in average air temperature [36]. Therefore, it is important to understand the phylogeography of L. dispar dispar in Europe regarding management strategies [37].
Major [38,39,40] and minor refugia [27,41] during ice ages coupled with recent gene flow events [42] determined the current pattern of divergence. Though for some species it has been demonstrated that the most recent ice age had the most profound impact on intraspecific diversity [43], for several other species the origin of divergence goes further back before the last ice age [44,45], with Lymantria dispar being one of them [13]. Intraspecific divergence among haplotypes of European L. dispar dispar (0.5029%) is congruent with previous findings that this clade arose around 200 thousand years (kyr) [13]. In particular, the observed genetic pattern of L. dispar dispar in Europe seems to have been shaped by at least two refugia, one located near the Carpathian Mountains (green) spreading to Central Europe and the Pontic-Mediterranean region (blue) spreading to the Balkans and the north of Europe. Furthermore, it is likely that contemporary pattern of gypsy moth phylogeography in Europe was also influenced by anthropogenic migration. This has been already hypothesized in the recent study of the Asian gypsy moth based on COI barcode sequences [46] that revealed an unusually close relationship between geographically distant haplotypes which could not be explained merely by natural dispersal of gypsy moth.
Being a highly polyphagous species, L. dispar populations would have been expected to exhibit deep intraspecific divergence as different selection pressures are associated with different host plants, something that favors the emergence of barriers to gene flow [47,48]. Should these barriers be maintained for a sufficiently long period of time, they may lead to the formation of distinct host-associated lineages [49]. This concept was initially thought to be valid mostly for parthenogenetically reproducing species [50,51,52], yet recent studies have shown that it is also evident in sexually reproducing species [49,53]. However, the limited genetic structure revealed among gypsy moth populations in Europe coupled with the AMOVA outcome do not support this expectation. As with other polyphagous insect species such as Pityogenes chalcographus [54] or Heliconius melpomene [55,56] life-history traits might have evolved on one host plant upon secondary contact could have been transmitted to populations feeding on a different host species, and thus render the herbivore insect species capable to feed on different host species [54]. Furthermore, this effect is reinforced by the high gene flow of L. dispar as the combination of the migratory pathways (flying moths, wind-borne dispersal of neonate larvae and long distances anthropogenic movement [9]) could facilitate the redistribution and coalescence of populations [57].

5. Conclusions

Despite the relatively large sample, we were able to determine only limited inferences of phylogeographic structure among localities in Europe, with at least two refugial areas contributing to the current genetic structure of gypsy moth. The near panmictic status of haplotype LD4 also present in Georgia suggests that more complex forces have taken role in shaping of the contemporary phylogeographic pattern of gypsy moth, and one of likely explanations could be a man-aided dispersal. Such hypotheses have already been drawn on larger scale studies [46]. However, the separation of the Georgian populations from other European populations, but also from the Asian haplotypes L. dispar asiatica and L. dispar japonica, shows that there are likely many more genotypic entities than recently described [28]. In fact, as differences in biology and behavior are described between gypsy moths from Europe and Asia, i.e., flight capability of females [10,58], the region of the Caucasus should be studied more thoroughly in order to examine the taxonomic status of the genetic entities in that geographic area.

Author Contributions

N.L. is a postdoc responsible to introduce and implement research of forest pests and diseases by means of molecular biology in Division of Forest Protection and Game Management at Croatian Forest Research Institute. The author team forms part of a consortium of a research project on defoliators as invasive forest pests in changing climate conditions (DIFPEST), coordinated by M.P. C.S., and M.P. conceived the ideas; N.L., D.N.A., and C.B. conducted experiments. N.L., C.B., and D.N.A. analyzed data. All authors contributed to the writing.

Funding

This research was partly funded by Croatian Science Foundation project HRZZ 7616 DIFPEST, and Bilateral scientific and technical cooperation project funded by Croatian Ministry of Science and ÖAD. This study was also supported by the Austrian Science Fund FWF project I2604-B25.

Acknowledgments

We thank Nicolas Meurisse, Jaroslav Holuša, Alain Roques, Horst Delb, Gabriela Lobinger, György Csóka, Milka Glavendekić, Ferenc Lakatos, Daniela Kirilova Pilarski, Leen Moraal, Pio Federico Roversi, Michele Squarcini, Gernot Hoch, Mario Contarini, Irena Papazova, Sterja Naceski, Constantin Netoiu, Medea Burjanadze, and Andreas Linde for generous help in sampling. We also thank to Susanne Kumböck, Hannes Schuler and Blaženka Ercegvac for help in the laboratory and Vassiliki Sgardeli, George I. Memtsas, and Nikolaos Fyllas for their assistance with R programming.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pogue, M.G.; Schaefer, P.W. A Review of Selected Species of Lymantria Hübner [1819] (Lepidoptera: Noctuidae: Lymantriinae) from Subtropical and Temperate Regions of Asia, Including Descriptions of Three New Species, Some Potentially Invasive to North America; U.S. Department of Agriculture, Forest Service, Forest Health Technology Enterprise Team: Washington, DC, USA, 2007; pp. 11–37.
  2. Schintlmeister, A. The taxonomy of the genus Lymantria Hübner, [1819] (Lepidoptera: Lymantriidae). Quadrifina 2004, 7, 1–248. [Google Scholar]
  3. Liebhold, A.M.; Mastro, V.C.; Schaefer, P.W. Learning from the legacy of Leopold Trouvelot. Bull. Entomol. Soc. Am. 1989, 35, 20–22. [Google Scholar] [CrossRef]
  4. McManus, M.L.; Csóka, G. History and Impact of Gypsy Moth in North America and Comparison to Recent Outbreaks in Europe. Acta Silv. Lignaria Hung. 2007, 3, 47–64. [Google Scholar]
  5. Ozumbekov, A.A.; Liebhold, A.M.; Ponomarev, V.I.; Tobin, P.C. Gypsy moth (Lepidoptera: Lymantriidae) in Central Asia. Am. Entomol. 2009, 55, 258–264. [Google Scholar] [CrossRef]
  6. Davidson Christopher, B.; Gottschalk Kurt, W.; Johnson James, E. European gypsy moth (Lymantria dispar L.) outbreaks: A review of the literature. In Gen Tech Rep NE-278; U.S. Department of Agriculture, Forest Service, Northeastern Research Station: Newtown Square, PA, USA, 2001; 15p. [Google Scholar]
  7. Cannon, R.J.C.; Koerper, D.; Ashby, S.; Baker, R.; Bartlett, P.W.; Brookes, G.; Burgess, R.; Cheek, S.; Evans, H.F.; Hammon, R.; et al. Gypsy moth, Lymantria dispar, outbreak in northeast London, 1995–2003. Int. J. Pest Manag. 2004, 50, 259–273. [Google Scholar] [CrossRef]
  8. Elkinton, J.S.; Liebhold, A.M. Population dynamics of gypsy moth in North America. Annu. Rev. Entomol. 1990, 35, 571–596. [Google Scholar] [CrossRef]
  9. Tobin, P.C.; Blackburn, L.M. Long-distance dispersal of the gypsy moth (Lepidoptera: Lymantridae) facilitated its initial invasion of Wisconsin. Environ. Entomol. 2008, 37, 87–93. [Google Scholar] [CrossRef]
  10. Keena, M.A.; Côte, M.J.; Grinberg, P.S.; Wallner, W.E. World distribution of female flight and genetic variation in Lymantria dispar (Lepidoptera: Lymantriidae). Environ. Entomol. 2008, 37, 636–649. [Google Scholar] [CrossRef]
  11. Bogdanowicz, S.M.; Wallner, W.E.; Bell, J.; O’Dell, T.M.; Harrison, R.G. Asian gypsy moth (Lepidoptera: Lymantriidae) in North America: Evidence from molecular data. Ann. Entomol. Soc. Am. 1993, 86, 710–715. [Google Scholar] [CrossRef]
  12. Kang, T.H.; Sang, H.H.; Heung, S.L. Genetic structure and demographic history of Lymantria dispar (Linnaeus, 1758) (Lepidoptera: Erebidae) in its area of origin and adjacent areas. Ecol. Evol. 2017, 7, 9162–9178. [Google Scholar] [CrossRef] [PubMed]
  13. Wu, Y.; Molongoski, J.J.; Winograd, D.F.; Bogdanowicz, S.M.; Louyakis, A.S.; Lance, D.R.; Mastro, V.C.; Harrison, R.G. Genetic structure, admixture and invasion success in a Holarctic defoliator, the gypsy moth (Lymantria dispar, Lepidoptera: Erebidae). Mol. Ecol. 2015, 24, 1275–1291. [Google Scholar] [CrossRef] [PubMed]
  14. deWaard, J.R.; Mitchell, A.; Keena, M.A.; Gopurenko, D.; Boykin, L.M.; Armstrong, K.F.; Pogue, M.G.; Lima, J.; Floyd, R.; Hanner, R.H.; et al. Towards a global barcode library for Lymantria (Lepidoptera: Lymantriinae) tussock moths of biosecurity concern. PLoS ONE 2010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Quian, L.; Yulin, A.; Juncian, S.; Mei, X.; Jianlin, Y.; Cuiping, W.; Bin, L.; Dejun, H. COI gene geographic variation of gypsy moth (Lepidoptera: Lymantriidae) and a TaqMan PCR diagnostic assay. DNA Barcodes 2014, 2, 10–16. [Google Scholar] [CrossRef]
  16. Lacković, N.; Bertheau, C.; Stauffer, C.; Pernek, M.; Avtzis, D.N. Genetic split between coastal and continental populations of gypsy moth separated by Dinaric Alps. J. Appl. Entomol. 2016, 139, 721–726. [Google Scholar] [CrossRef]
  17. Lunt, D.H.; Zhang, D.X.; Szymura, J.M.; Hewitt, G.M. The insect cytochrome oxidase I gene: Evolutionary patterns and conserved primers for phylogenetic studies. Insect Mol. Biol. 1996, 5, 153–165. [Google Scholar] [CrossRef] [PubMed]
  18. Hall, T.A. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucl. Acid. Res. 1999, 41, 95–98. [Google Scholar]
  19. Excoffier, L.; Lischer, H.E.L. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour. 2010, 10, 564–567. [Google Scholar] [CrossRef] [PubMed]
  20. Petit, R.J.; El Mousadik, A.; Pons, O. Identifying populations for conservation on the basis of genetic markers. Conserv. Biol. 1998, 12, 844–855. [Google Scholar] [CrossRef]
  21. Clement, M.; Posada, D.; Crandall, K.A. TCS: A computer program to estimate gene genealogies. Mol. Ecol. 2000, 9, 1657–1659. [Google Scholar] [CrossRef] [PubMed]
  22. Crandall, K.A.; Templeton, A.R. Empirical tests of some predictions from coalescent theory with applications to intraspecific phylogeny reconstruction. Genetics 1993, 135, 959–969. [Google Scholar]
  23. Panchal, M. The automation of Nested Clade Phylogeographic Analysis. Bioinformatics 2007, 23, 509–510. [Google Scholar] [CrossRef] [PubMed]
  24. Pons, O.; Petit, R.J. Measuring and testing genetic differentiation with ordered versus unordered alleles. Genetics 1996, 144, 1237–1245. [Google Scholar] [PubMed]
  25. Darriba, D.; Taboada, G.L.; Doallo, R.; Posada, D. jModelTest 2: More models, new heuristics and parallel computing. Nat. Methods 2012, 9, 772. [Google Scholar] [CrossRef] [PubMed]
  26. Ronquist, F.; Teslenko, M.; Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic interface and model selection across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [PubMed]
  27. Stewart, D.; Zahiri, R.; Djoumad, A.; Freschi, L.; Lamarche, J.; Holden, D.; Cervantes, S.; Ojeda, D.I.; Potvin, A.; Nisole, A.; et al. A Multi-Species TaqMan PCR Assay for the Identification of Asian Gypsy Moths (Lymantria spp.) and Other Invasive Lymantriines of Biosecurity Concern to North America. PLoS ONE 2016, 11. [Google Scholar] [CrossRef] [PubMed]
  28. Djoumad, A.; Nisole, A.; Zahiri, R. Comparative analysis of mitochondrial genomes of geographic variants of the gypsy moth, Lymantria dispar, reveals a previously undescribed genotypic entity. Sci. Rep. 2017, 7, 142–145. [Google Scholar] [CrossRef] [PubMed]
  29. Vandergast, A.G.; Perry, W.M.; Lugo, R.V.; Hathaway, S.A. Genetic landscapes GIS Toolbox: Tools to map patterns of genetic divergence and diversity. Mol. Ecol. Resour. 2011, 11, 158–161. [Google Scholar] [CrossRef] [PubMed]
  30. R Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. 2015. Available online: https://www.R-project.org/ (accessed on 20 September 2018).
  31. Oksanen, J.; Guillaume Blanchet, F.; Kindt, R.; Legendre, P.; Minchin, P.R.; O’Hara, R.B.; Simpson, G.L.; Solymos, P.; Stevens, M.H.H.; Szoecs, E.; et al. Vegan: Community Ecology Package. R Package Version 2.3-3. 2016. Available online: https://CRAN.Rproject.org/package=vegan (accessed on 20 September 2018).
  32. Phillips, J.D.; Gwiazdowski, R.A.; Ashlock, D.; Hanner, R. An exploration of sufficient sampling effort to describe intraspecific DNA barcode haplotype diversity: Examples from the ray-finned fishes (Chordata: Actinopterygii). DNA Barcodes 2015, 3, 66–73. [Google Scholar] [CrossRef]
  33. Biogeographic Regions in Europe. Available online: https://www.eea.europa.eu/data-and-maps/figures/biogeographical-regions-in-europe-2 (accessed on 19 January 2016).
  34. Karolewski, P.; Grzebyta, J.; Oleksyn, J.; Giertych, M.J. Effects of temperature on larval survival rate and duration of development in Lymantria monacha (L.) on needles of Pinus sylvestris (L.) and L. dispar (L.) on leaves of Quercus robur (L.). Pol. J. Ecol. 2007, 55, 595–600. [Google Scholar]
  35. Tobin, P.C.; Liebhold, A.M. Gypsy moth. In Encyclopedia of Biological Invasions; Simberloff, D., Rejmanek, M., Eds.; University of California Press: Berkeley, CA, USA, 2011; pp. 298–304. [Google Scholar]
  36. Vanhanen, H.; Veteli, T.O.; Päivinen, S.; Kellomäki, S.; Niemelä, P. Climate Change and Range Shifts in Two Insect Defoliators: Gypsy Moth and Nun Moth—A Model Study. Silva Fenn. 2007, 41, 621–638. [Google Scholar] [CrossRef]
  37. Ochando, M.D.; Reyes, A.; Segura, D.; Callejas, D. Phylogeography: It’s Importance in Insect Pest Control; Nova Science Publishers Inc.: New York, NY, USA, 2011; p. 48. [Google Scholar]
  38. Taberlet, P.; Fumagalli, L.; Wust-Saucy, A.G.; Cosson, J.F. Comparative phylogeography and postglacial colonization routes in Europe. Mol. Ecol. 1998, 7, 453–464. [Google Scholar] [CrossRef] [PubMed]
  39. Hewitt, G.M. Post-glacial re-colonization of European biota. Biol. J. Linn. Soc. 1999, 68, 87–112. [Google Scholar] [CrossRef] [Green Version]
  40. Hewitt, G.M. The genetic legacy of the Quaternary ice ages. Nature 2000, 405, 907–913. [Google Scholar] [CrossRef] [PubMed]
  41. Stewart, J.R.; Lister, A.M. Cryptic northern refugia and the origins of modern biota. Trends Ecol. Evol. 2001, 16, 608–613. [Google Scholar] [CrossRef]
  42. Excoffier, L.; Foll, M.; Petit, R.J. Genetic consequences of range expansions. Annu. Rev. Ecol. Evol. Syst. 2009, 40, 481–501. [Google Scholar] [CrossRef]
  43. Schmitt, T. Molecular biogeography of Europe: Pleistocene cycles and postglacial trends. Front. Zool. 2007, 4, 11. [Google Scholar] [CrossRef] [PubMed]
  44. Haubrock, P.J.; Kvifte, G.M.; Langguth, H.; Wagner, R. Glacial and postglacial species divergence and dispersal of European trickle midges (Diptera: Thaumaleidae). Arthropod Syst. Phyl. 2017, 75, 523–534. [Google Scholar]
  45. Tóth, V.; Lakatos, F. Phylogeographic pattern of the plane leaf miner, Phyllonorycter platani (Staudinger, 1870) (Lepidoptera: Gracilariidae) in Europe. BMC Evol. Biol. 2018, 18, 135. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, F.; Luo, Y.; Keena, M.A.; Wu, Y.; Shi, J. DNA barcoding of Chinese gypsy moths (Lepidoptera, Erebidae) reveals new haplotypes and divergence patterns within gypsy moth subspecies. J. Econ. Entomol. 2016, 109, 366–374. [Google Scholar] [CrossRef] [PubMed]
  47. Feder, J.L.; Hunt, T.A.; Bush, L. The effects of climate, host plant phenology and host fidelity on the genetics of apple and hawthorn infesting races of Rhagoletis pomonella. Entomol. Exp. Appl. 1993, 69, 117–135. [Google Scholar] [CrossRef]
  48. Nosil, P.; Crespi, B.J. Ecological divergence promotes the evolution of cryptic reproductive isolation. P. Roy. Soc. B-Biol. Sci. 2006, 273, 991–997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Antwi, J.B.; Sword, G.A.; Medina, R.F. Host-associated differentiation in a highly polyphagous, sexually reproducing insect herbivore. Ecol. Evol. 2015, 5, 2533–2543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Peccoud, J.; Ollivier, A.; Plantegenest, M.; Simon, J.C. A continuum of genetic divergence from sympatric host races to species in the pea aphid complex. Proc. Natl. Acad. Sci. USA 2009, 106, 7495–7500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Via, S. Natural selection in action during speciation. Proc. Natl. Acad. Sci. USA 2009, 106, 9939–9946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Marques, J.F.; Wang, H.L.; Svensson, G.P.; Frago, E.; Anderbrant, O. Genetic divergence and evidence for sympatric host-races in the highly polyphagous brown tail moth, Euproctis chrysorrhoea (Lepidoptera: Erebidae). Evol. Ecol. 2014, 28, 829–848. [Google Scholar] [CrossRef]
  53. Hereward, J.P.; Walter, G.H.; DeBarro, P.J.; Lowe, A.J.; Riginos, C. Gene flow in the green mirid, Creontiades dilutus (Hemiptera: Miridae), across arid and agricultural environments with different host plant species. Ecol. Evol. 2013, 3, 807–821. [Google Scholar] [CrossRef] [PubMed]
  54. Bertheau, C.; Bankhead-Dronnet, S.; Martin, C.; Lieutier, F.; Roux-Morabito, G. Lack of genetic differentiation after host range extension argues for the generalist nature of Pityogenes chalcographus (Curculionidae: Scolytinae). Ann. For. Sci. 2012, 69, 313–323. [Google Scholar] [CrossRef]
  55. Dasmahapatra, K.K.; Walters, J.R.; Briscoe, A.D.; Davey, J.W.; Whibley, A.; Nadeau, N.J.; Zimin, A.V.; Hughes, D.S.T.; Ferguson, L.C.; Martin, S.H.; et al. Butterfly genome reveals promiscuous exchange of mimicry adaptations among species. Nature 2012, 487, 94–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Nadeau, N.J.; Martin, S.H.; Kozak, K.M.; Salazar, C.; Dasmahapatra, K.K.; Davey, J.W.; Baxter, S.W.; Blaxter, M.L.; Mallet, J.; Jiggins, C.D. Genome-wide patterns of divergence and gene flow across a butterfly radiation. Mol. Ecol. 2013, 22, 814–826. [Google Scholar] [CrossRef] [PubMed]
  57. McFadden, M.W.; McManus, M.L. An Insect Out of Control? The potential for spread and establishment of the gypsy moth in new forest areas in the United States. In Forest Insect Guilds: Patterns of Interaction with Host Trees; Baranchikov, Y.N., Mattson, W.J., Hain, F.P., Payne, T.L., Eds.; U.S. Department of Agriculture, Forest Service: Radnor, PA, USA, 1991; pp. 172–186. [Google Scholar]
  58. Keena, M.A.; Grinberg, P.S.; Wallner, W.E. Inheritance of female flight in Lymantria dispar (Lepidoptera: Lymantriidae). Environ. Entomol. 2007, 36, 484–494. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Parsimony network reconstructed by cladistic analysis with 95% connection limit using TCS 1.21. Numbers represent the haplotypes. Dimensions of pies are proportional to frequency, and each haplotype is colored based on its affiliation to a certain clade.
Figure 1. Parsimony network reconstructed by cladistic analysis with 95% connection limit using TCS 1.21. Numbers represent the haplotypes. Dimensions of pies are proportional to frequency, and each haplotype is colored based on its affiliation to a certain clade.
Insects 09 00143 g001
Figure 2. Haplotype distribution of Lymantria dispar in Europe. Colored circles represent the clades according to parsimony network (Figure 1), acronyms indicate localities (Table 1) and dimension of pies are proportional to sample size and haplotype abundance. Underlaying map represents biogeographic regions of Europe [33]. The figure was constructed using ArcMap 10.1 (ESRI, Redlands, CA, USA).
Figure 2. Haplotype distribution of Lymantria dispar in Europe. Colored circles represent the clades according to parsimony network (Figure 1), acronyms indicate localities (Table 1) and dimension of pies are proportional to sample size and haplotype abundance. Underlaying map represents biogeographic regions of Europe [33]. The figure was constructed using ArcMap 10.1 (ESRI, Redlands, CA, USA).
Insects 09 00143 g002
Figure 3. Bayesian reconstruction of molecular phylogenetic relationships among haplotypes of the current research supplemented with sequences submitted in GenBank (see text for detailed description). Numbers above nodes represent the Bayesian posterior probability with values greater than 0.6. Color of each bar is identical to the ones used in Figure 1.
Figure 3. Bayesian reconstruction of molecular phylogenetic relationships among haplotypes of the current research supplemented with sequences submitted in GenBank (see text for detailed description). Numbers above nodes represent the Bayesian posterior probability with values greater than 0.6. Color of each bar is identical to the ones used in Figure 1.
Insects 09 00143 g003
Table 1. Countries, localities, WGS84 coordinates in degrees (Lat., Long.), host tree (QE = Quercus petraea; QR = Q. robur; QI = Q. ilex; QP = Q. pubescens; QS = Q. suber; QO = Q. coccifera; FS = Fagus sylvatica; MA = Malus sp.; VAR1 = Q.petraea + F.sylvatica; VAR2 = Q. robur + F. sylvatica; VAR3 = Q. robur + Q. petraea; VAR4 = Q. petraea + Q. pubescens; VAR5 = Salix sp. + Populus sp.; VAR6 = Q. rubra + Q. petraea + Q. cerris + Carpinus betulus + Fraxinus sp.), number of individuals (N), number of haplotypes (HT), haplotype diversity (Hd), nucleotide diversity (π), mean number of pairwise differences (MNPD), and Allelic richness (r) after rarefaction to smallest sample size (5). Mean value and standard deviation (SD) is shown for all genetic diversity indices except for r.
Table 1. Countries, localities, WGS84 coordinates in degrees (Lat., Long.), host tree (QE = Quercus petraea; QR = Q. robur; QI = Q. ilex; QP = Q. pubescens; QS = Q. suber; QO = Q. coccifera; FS = Fagus sylvatica; MA = Malus sp.; VAR1 = Q.petraea + F.sylvatica; VAR2 = Q. robur + F. sylvatica; VAR3 = Q. robur + Q. petraea; VAR4 = Q. petraea + Q. pubescens; VAR5 = Salix sp. + Populus sp.; VAR6 = Q. rubra + Q. petraea + Q. cerris + Carpinus betulus + Fraxinus sp.), number of individuals (N), number of haplotypes (HT), haplotype diversity (Hd), nucleotide diversity (π), mean number of pairwise differences (MNPD), and Allelic richness (r) after rarefaction to smallest sample size (5). Mean value and standard deviation (SD) is shown for all genetic diversity indices except for r.
CountryAbbrev.HostLat.Long.NHTHd ± SDπ ± SDMNPD ± SDr (5)
AustriaAUT1QE47.7516.542750.695 ± 0.0780.001 ± 0.0010.871 ± 0.6321.966
BelgiumBEL1FS50.824.411120.181 ± 0.1430.001 ± 0.0000.727 ± 0.5840.455
BulgariaBUL1QR43.0923.46510 ± 00 ± 00 ± 00.000
BUL2QR41.7124.16830.607 ± 0.1640.002 ± 0.0011.357 ± 0.9331.518
BUL3QR42.7427.74510 ± 00 ± 00 ± 00.000
BUL4QR43.2126.621230.727 ± 0.0580.002 ± 0.0011.939 ± 1.1811.788
CroatiaCRO1QI44.6514.4737100.818 ± 0.0380.003 ± 0.0022.087 ± 1.1932.555
CRO2QR45.4717.991410 ± 00 ± 00 ± 00.000
CRO3QR44.9519.07510 ± 00 ± 00 ± 00.000
CRO4QP44.1515.31980.859 ± 0.0520.004 ± 0.0022.666 ± 1.4852.815
CRO5VA145.2216.282560.426 ± 0.1210.000 ± 0.0000.473 ± 0.4251.167
CRO6QR45.6515.59620.333 ± 0.2150.000 ± 0.0000.333 ± 0.3800.833
Czech RepublicCZE1QE49.1117.041040.533 ± 0.1800.001 ± 0.0010.600 ± 0.5191.500
CZE2QR48.8817.1720.285 ± 0.1960.000 ± 0.0000.285 ± 0.3400.714
CZE3QE48.8516.01730.523 ± 0.2080.001 ± 0.0010.857 ± 0.6811.429
FranceFRA1QS41.719.331130.345 ± 0.1720.000 ± 0.0000.363 ± 0.3770.909
FRA2QP44.472.47520.400 ± 0.2370.000 ± 0.0000.400 ± 0.4351.000
FRA3QE47.821.922470.721 ± 0.0700.002 ± 0.0011.554 ± 0.9612.038
FRA4QE44.852.121460.802 ± 0.0900.001 ± 0.0011.219 ± 0.8242.500
GeorgiaGEO1MA41.9444.481130.618 ± 0.1030.008 ± 0.0055.818 ± 3.0131.407
GEO2MA41.8846.13630.733 ± 0.1550.005 ± 0.0033.933 ± 2.2901.833
GermanyGER1VA248.267.762060.636 ± 0.1150.001 ± 0.0011.194 ± 0.7971.838
GER2VA349.9110.192660.566 ± 0.1080.001 ± 0.0011.141 ± 0.7641.579
GER3QE51.9114.35810 ± 00 ± 00 ± 00.000
GreeceGRE1QO40.7723.113130.127 ± 0.0790.000 ± 0.0000.193 ± 0.2500.323
GRE2QO41.0823.541220.303 ± 0.1470.000 ± 0.0000.303 ± 0.3360.682
GRE3QO38.6322.97510 ± 00 ± 00 ± 00.000
GRE4QO41.1225.41920.500 ± 0.1280.001 ± 0.0011.000 ± 0.7390.952
HollandHOL1QR51.985.68930.555 ± 0.1650.000 ± 0.0000.611 ± 0.5301.389
HungaryHUN1VA447.8319.961040.533 ± 0.1800.001 ± 0.0010.800 ± 0.6281.500
HUN2VA547.7117.121140.709 ± 0.0990.001 ± 0.0010.872 ± 0.6611.851
HUN3VA647.0717.34520.600 ± 0.1750.000 ± 0.0000.600 ± 0.5621.000
ItalyITA1QI42.9910.52550.776 ± 0.0460.002 ± 0.0011.513 ± 0.9412.245
ITA2QS40.518.461430.483 ± 0.1420.000 ± 0.0000.527 ± 0.4671.209
FYROMMAC1N/A41.3621.21740.809 ± 0.1290.002 ± 0.0011.714 ± 1.1312.381
PolandPOL1VA553.2922.611530.361 ± 0.1440.001 ± 0.0010.590 ± 0.5000.905
RomaniaROM1N/A44.523.741410 ± 00 ± 00 ± 00.000
SerbiaSER1QE44.475.57730.523 ± 0.2080.001 ± 0.0010.857 ± 0.6811.429

Share and Cite

MDPI and ACS Style

Lacković, N.; Pernek, M.; Bertheau, C.; Franjević, D.; Stauffer, C.; Avtzis, D.N. Limited Genetic Structure of Gypsy Moth Populations Reflecting a Recent History in Europe. Insects 2018, 9, 143. https://doi.org/10.3390/insects9040143

AMA Style

Lacković N, Pernek M, Bertheau C, Franjević D, Stauffer C, Avtzis DN. Limited Genetic Structure of Gypsy Moth Populations Reflecting a Recent History in Europe. Insects. 2018; 9(4):143. https://doi.org/10.3390/insects9040143

Chicago/Turabian Style

Lacković, Nikola, Milan Pernek, Coralie Bertheau, Damjan Franjević, Christian Stauffer, and Dimitrios N. Avtzis. 2018. "Limited Genetic Structure of Gypsy Moth Populations Reflecting a Recent History in Europe" Insects 9, no. 4: 143. https://doi.org/10.3390/insects9040143

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop