global phylogeography and geographical variation in ...users.jyu.fi/~mappes/publications/hegna 2015-...

13
ORIGINAL ARTICLE Global phylogeography and geographical variation in warning coloration of the wood tiger moth (Parasemia plantaginis) Robert H. Hegna 1,2 *, Juan A. Galarza 1 and Johanna Mappes 1 1 University of Jyvaskyla, Centre of Excellence in Biological Interactions, Department of Biological and Environmental Science, 40014 Jyvaskyla, Finland, 2 Department of Biology, Palm Beach Atlantic University, West Palm Beach, FL 33401, USA *Correspondence: Robert H. Hegna, Department of Biology, Palm Beach Atlantic University, 901 S. Flagler Dr., West Palm Beach, FL 33401, USA. E-mail: [email protected] ABSTRACT Aim To investigate the phylogeography of the aposematic wood tiger moth (Parasemia plantaginis) across its Holarctic distribution and to explore how its genetic structure relates to geographical differences in hindwing warning color- ation of males and females. Males have polymorphic hindwing coloration, while female hindwing coloration varies continuously, but no geographical analyses of coloration or genetic structure exist. Location The Holarctic. Methods We sequenced a fragment of the mitochondrial cytochrome c oxidase subunit I gene (COI) from 587 specimens. We also examined more current population structure by genotyping 569 specimens at 10 nuclear micro- satellite loci. Species distribution modelling for present conditions and the Last Glacial Maximum (LGM) was performed to help understand genetic structure. Geographical patterns in hindwing warning coloration were described from 1428 specimens and compared to the genetic analyses. Results We found only two instances of genetic divergence that coincided with distinct, yet imperfect, shifts in male hindwing coloration in the Caucasus region and Japan. A shift in female hindwing colour did not appear to be asso- ciated with genetic structure. A change from sexual monomorphism to sexual dimorphism was also observed. Mitogenetic (mtDNA) structure does not show the influence of glacial refugia during the LGM. Climate shifts following the LGM appear to have isolated the red Caucasus populations and other southerly populations. Populations at opposite ends of the moth’s distribution showed high levels of differentiation in the microsatellite data analysis compared to the shallow mitogenetic structure, supporting a more recent divergence. Main conclusions Parasemia plantaginis populations appeared to have been historically well connected, but current populations are much more differenti- ated. This raises the possibility that incipient speciation may be occurring in portions of the species’ distribution. Some changes in colour align to genetic differences, but others do not, which suggests a role for selective and non- selection based influences on warning signal variation. Keywords Aposematism, Arctiidae, Arctiinae, colour polymorphism, Erebidae, Holarctic, Lepidoptera, sexual dimorphism, species distribution model. INTRODUCTION Studies of phenotypic differences across species’ distributions provide valuable insight into the processes of natural selection and speciation (Endler, 1977). While many species can exhibit varying degrees of differentiation across their distributions (Endler, 1977), aposematic organisms with warning signals that advertise anti-predator defences offer a ª 2015 John Wiley & Sons Ltd http://wileyonlinelibrary.com/journal/jbi 1 doi:10.1111/jbi.12513 Journal of Biogeography (J. Biogeogr.) (2015)

Upload: others

Post on 02-Feb-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

  • ORIGINALARTICLE

    Global phylogeography and geographicalvariation in warning coloration ofthe wood tiger moth (Parasemiaplantaginis)Robert H. Hegna1,2*, Juan A. Galarza1 and Johanna Mappes1

    1University of Jyvaskyla, Centre of Excellence

    in Biological Interactions, Department of

    Biological and Environmental Science, 40014

    Jyv€askyl€a, Finland, 2Department of Biology,

    Palm Beach Atlantic University, West Palm

    Beach, FL 33401, USA

    *Correspondence: Robert H. Hegna,

    Department of Biology, Palm Beach Atlantic

    University, 901 S. Flagler Dr., West Palm

    Beach, FL 33401, USA.

    E-mail: [email protected]

    ABSTRACT

    Aim To investigate the phylogeography of the aposematic wood tiger moth

    (Parasemia plantaginis) across its Holarctic distribution and to explore how its

    genetic structure relates to geographical differences in hindwing warning color-

    ation of males and females. Males have polymorphic hindwing coloration,

    while female hindwing coloration varies continuously, but no geographical

    analyses of coloration or genetic structure exist.

    Location The Holarctic.

    Methods We sequenced a fragment of the mitochondrial cytochrome c

    oxidase subunit I gene (COI) from 587 specimens. We also examined more

    current population structure by genotyping 569 specimens at 10 nuclear micro-

    satellite loci. Species distribution modelling for present conditions and the Last

    Glacial Maximum (LGM) was performed to help understand genetic structure.

    Geographical patterns in hindwing warning coloration were described from

    1428 specimens and compared to the genetic analyses.

    Results We found only two instances of genetic divergence that coincided

    with distinct, yet imperfect, shifts in male hindwing coloration in the Caucasus

    region and Japan. A shift in female hindwing colour did not appear to be asso-

    ciated with genetic structure. A change from sexual monomorphism to sexual

    dimorphism was also observed. Mitogenetic (mtDNA) structure does not show

    the influence of glacial refugia during the LGM. Climate shifts following the

    LGM appear to have isolated the red Caucasus populations and other southerly

    populations. Populations at opposite ends of the moth’s distribution showed

    high levels of differentiation in the microsatellite data analysis compared to the

    shallow mitogenetic structure, supporting a more recent divergence.

    Main conclusions Parasemia plantaginis populations appeared to have been

    historically well connected, but current populations are much more differenti-

    ated. This raises the possibility that incipient speciation may be occurring in

    portions of the species’ distribution. Some changes in colour align to genetic

    differences, but others do not, which suggests a role for selective and non-

    selection based influences on warning signal variation.

    Keywords

    Aposematism, Arctiidae, Arctiinae, colour polymorphism, Erebidae, Holarctic,

    Lepidoptera, sexual dimorphism, species distribution model.

    INTRODUCTION

    Studies of phenotypic differences across species’ distributions

    provide valuable insight into the processes of natural

    selection and speciation (Endler, 1977). While many species

    can exhibit varying degrees of differentiation across their

    distributions (Endler, 1977), aposematic organisms with

    warning signals that advertise anti-predator defences offer a

    ª 2015 John Wiley & Sons Ltd http://wileyonlinelibrary.com/journal/jbi 1doi:10.1111/jbi.12513

    Journal of Biogeography (J. Biogeogr.) (2015)

  • unique perspective on such variation because they are pre-

    dicted to be under stabilizing selection (Joron & Mallet,

    1998). Geographical variation in warning signals is often

    found to be the result of co-evolution between the apose-

    matic prey and local predators (Kapan, 2001; Chouteau &

    Angers, 2012). Thus, studying geographical variation in

    warning signals in relation to genetic variation increases our

    understanding about warning signal evolution and diversifi-

    cation (H€arlin & H€arlin, 2003; Wang & Shaffer, 2008; Hines

    et al., 2011). However, many species with geographically

    varying warning signals studied to date occur in tropical eco-

    systems, where predator/prey relationships are thought to be

    more tightly intertwined (Schemske et al., 2009). Therefore,

    it is also necessary to study aposematic species distributed in

    the north that have experienced different climatic histories

    and ecological conditions to fully understand how intraspe-

    cific warning signal variation evolves.

    We investigated the phylogeography, population structure,

    and hindwing warning signal polymorphism in the apose-

    matic wood tiger moth, Parasemia plantaginis (Linnaeus,

    1758), across its Holarctic distribution to explore evolution-

    ary influences on warning coloration. Hindwing colour func-

    tions as a warning signal in males and females (Nokelainen

    et al., 2012), and is heritable (Nokelainen et al., 2013).

    Variable melanized markings on the hindwings can obscure

    the colour completely in some areas and limit warning signal

    efficacy (Hegna et al., 2013). Male hindwing colour is known

    to be polymorphic over much of its distribution, but no

    study has systematically examined the distribution or fre-

    quency of colours across the entire distribution.

    In the Caucasus region, male hindwing coloration is

    described as reddish (Koc�ak, 1989; De Freina, 1993; Dubato-lov & Zahiri, 2005). This stands in contrast to neighbouring

    populations that have polymorphic males with either yellow

    or white hindwings. Such divergence in coloration raises

    questions as to how closely connected populations in this

    region are to adjacent European and south-western Russian

    populations. Given the large species distribution (c.

    17,000 km), additional changes in hindwing colour may

    occur that are not yet described. For instance, some work

    has suggested that Japanese populations mostly comprise

    males with white hindwings (Okano & Katayama, 1976;

    Kishida, 2011), and variation in coloration is also reported

    to occur within North America (Coolidge, 1910). However,

    none of these descriptions follow from an examination of

    large numbers of specimens and almost nothing is known

    about geographical variation in female warning coloration.

    Here, we describe variation in hindwing colour across the

    distribution of P. plantaginis. We then evaluate the historical

    and recent genetic structure of the species across its Holarc-

    tic distribution using two independent genetic markers and

    determine whether genetic structuring matches any shifts in

    male or female hindwing colour. Alignment between hind-

    wing colour and genetic differences in neutral markers could

    suggest that geographical isolation and drift are primary

    drivers of hindwing colour divergence, rather than prezygotic

    isolation. Genetic and coloration differences could be either

    the product of isolation as a result of climate and/or geo-

    graphical factors, or the result of local selection by predators

    as observed in other aposematic species. We employed cli-

    mate-based species distribution modelling to explore whether

    climatic isolation could be a factor in genetic and coloration

    differences. In examining hindwing and genetic data concur-

    rently we aimed to develop a hypothesis about P. plantaginis

    origins that would help in understanding hindwing colour

    divergence in this species. Because hindwing coloration in

    the Caucasus region clearly differs from that in the sur-

    rounding areas (De Freina, 1993; Dubatolov & Zahiri, 2005),

    we hypothesized that the Caucasus region would be the most

    genetically diverged. We also hypothesized that changes in

    coloration would be imperfectly matched to genetic differ-

    ences in neutral markers, suggesting that processes other

    than drift and mutation (i.e. selective pressures) could be

    involved in hindwing colour evolution.

    MATERIALS AND METHODS

    Specimen collection for genetic analyses

    We obtained adult specimens of Parasemia plantaginis from

    across the entire Holarctic distribution (see Appendix S1a)

    through fieldwork, private collectors and museum collec-

    tions. Moths were captured in the field with butterfly nets

    and pheromone traps baited with calling females, which

    assisted in catching males. Methods for DNA extraction fol-

    lowed those of Galarza et al. (2010), with the exception that

    we extracted DNA from two legs rather than from the brain.

    Mitochondrial DNA genotyping and analysis

    We amplified a 630 bp fragment of the mitochondrial cyto-

    chrome c oxidase subunit I gene (COI) from 587 specimens

    (GenBank accession numbers: KP217220–KP217806; seeAppendix S1b). We used the COI sequences to reconstruct a

    maximum parsimony haplotype network to depict relation-

    ships among haplotypes of all 587 individuals using tcs

    1.2.1, setting a 95% connection limit (Clement et al., 2000).

    To further investigate relationships among haplotypes, we

    employed a Bayesian approach using MrBayes 3.2.1

    (Ronquist & Huelsenbeck, 2003). Our analysis included rep-

    resentatives of each haplotype and a representative from each

    geographical region bearing each haplotype. In total, we used

    82 specimens in the MrBayes analysis. The HKY+G modelwas the most appropriate substitution model indicated by

    jModelTest2 (Darriba et al., 2012) using both Akaike infor-

    mation criterion corrected for small sample size (AICc) and

    Bayesian information criterion (BIC). In the Bayesian

    analysis, two runs of four Markov chain Monte Carlo

    (MCMC) chains were run simultaneously (one cold and

    three hot) for 20 million generations. We set the tree

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    2

    R. H. Hegna et al.

    http://www.ncbi.nlm.nih.gov/nuccore/KP217220http://www.ncbi.nlm.nih.gov/nuccore/KP217806https://www.researchgate.net/publication/12275236_Clement_MD_Posada_D_Crandall_KA_TCS_a_computer_program_to_estimate_gene_genealogies_Mol_Ecol_9_1657-1659?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/27371926_Western_Lepidoptera-III-Notes_on_Leptarctia_Californiae_Walker?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/225896388_First_microsatellite_panel_for_the_Wood_Tiger_Moth_Parasemia_plantaginis?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/235384852_To_quiver_or_to_shiver_Increased_melanization_benefits_thermoregulation_but_reduces_warning_signal_efficacy_in_the_wood_tiger_moth?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/235378173_Environment-mediated_morph-linked_immune_and_life-history_responses_in_the_aposematic_wood_tiger_moth?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/10618138_MRBAYES_3_Bayesian_Phylogenetic_inference_under_mixed_models?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==

  • sampling rate to be once every 2000 generations. Following

    the completion of the analysis the relative burn-in rate was

    set to 0.25. Convergence was checked using Tracer 1.6

    (Rambaut & Drummond, 2003). The remaining trees were

    used to determine a majority-rule consensus tree.

    In order to infer demographic history we used DnaSP

    5.10 (Librado & Rozas, 2009) to calculate haplotype diversity

    (h), nucleotide diversity (p), Tajima’s D, Fu’s FS and averagenucleotide differences.

    Genotyping and analysis of nuclear microsatellite

    markers

    We amplified a total of 10 nuclear microsatellite loci as iden-

    tified by Galarza et al. (2010): Ppla107, Ppla109, Ppla279,

    Ppla313, Ppla317, Ppla323, Ppla363, Ppla382, Ppla414 and

    Ppla439. A total of 569 specimens were included in the final

    analyses because some older specimens failed to amplify for

    a majority of loci (see Appendices S1a, S1c & S2a).

    We used structure 2.3.4 (Pritchard et al., 2000) to esti-

    mate the number of populations (K) in our dataset that are

    relevant for studying evolutionary processes (see Appendix

    S1d). From the populations identified with structure, we

    constructed a neighbour-joining tree with 1000 bootstrap rep-

    licates using the DA measure of genetic distance (Nei et al.,

    1983) in poptree2 (Takezaki et al., 2010). We conducted a

    hierarchical analysis of molecular variance (AMOVA) using

    genalex 6.5 (Peakall & Smouse, 2006) to examine variation

    within populations, among six regions, and among popula-

    tions within the regions. The six regions were based on the

    relationships among populations in the poptree2 analysis

    and included the Caucasus, Europe, Russia, Japan, Alaska and

    the rest of North America (Canada–Utah). Alaskan popula-tions were included as a separate region in the AMOVA

    because the populations formed their own well-supported

    group in the poptree2 analysis. Russia comprised its own

    region in the AMOVA based on geography and no evidence

    of connectedness with Japanese populations, rather than by

    the results of the poptree2 analysis alone. We estimated FST(Weir, 1996) values among populations using the ENA

    method implemented in the program freena (Chapuis &

    Estoup, 2007) to account for any effects of null alleles, which

    can inflate FST values. We also calculated DEST (Jost, 2008)

    among all populations to supplement FST and DA measures of

    differentiation using genalex (see Appendix S2b).

    To investigate evidence of recent migration among popu-

    lations determined from our structure analysis we used the

    program BayesAss 3.0 (Wilson & Rannala, 2003). BayesAss

    implements a Bayesian method to estimate recent migration

    in the past three to four generations using MCMC sampling

    techniques (see Appendix S1e).

    Phenotypic data

    We used 1187 male and 241 female individuals to determine

    the distribution of hindwing warning signal colour across the

    Holarctic distribution of P. plantaginis (Table 1; also see

    Appendix S1a for detailed coordinates). Because the speci-

    mens were gathered from a total of 351 sampling sites (with

    some sites only having 1–5 moths) it was necessary to poolsamples collected from different sites in looking at the distri-

    bution of male and female hindwing colour across the Hol-

    arctic (Fig. 1c, see Appendix S1f for details). We visually

    determined the hindwing colour of males (white and yellow)

    because it is a polymorphic trait (white or yellow) without

    intermediate phenotypes (Galarza et al., 2014). Males from

    the Caucasus region with reddish hindwings were classified

    as red for simplicity on a large spatial scale and because we

    found no red individuals outside this area. Female hindwing

    colour varies continuously between red and yellow (Fig. 1),

    but can be matched to six standard colour tiles that were

    previously tested against the wing colour using spectropho-

    tometer data (see Lindstedt et al., 2011 for additional

    method detail). This visual classification method for female

    colour allowed us to reliably analyse geographical patterns

    without touching the fragile museum specimens with a spec-

    trophotometer. The number of specimens of each hindwing

    colour was counted for each region and analysed in ArcGIS

    10 (ESRI, Redlands, CA, USA).

    Table 1 Number of specimens obtained from differentcountries used to examine geographical variation in colorationacross the distribution of Parasemia plantaginis.

    Country Males Females Collector Field Museum

    Armenia 6 0 1 0 5

    Austria 96 8 4 100 0

    Azerbaijan 1 0 0 0 1

    Bulgaria 14 11 0 0 25

    Canada 110 11 3 108 10

    China 6 1 0 0 8

    Czech Republic 3 1 4 0 0

    Denmark 11 0 0 0 11

    Estonia 39 12 0 51 0

    Finland 246 41 0 287 0

    Georgia 14 6 0 0 20

    Italy 6 0 0 6 0

    Japan 92 11 0 103 0

    Kazakhstan 32 3 0 0 35

    Latvia 2 0 0 0 2

    Macedonia 1 1 2 0 0

    Mongolia 17 8 0 0 25

    North Korea 2 0 0 0 2

    Norway 29 0 0 0 29

    Poland 6 1 0 0 7

    Romania 2 0 2 0 0

    Russia 213 88 35 0 266

    Spain 1 0 0 0 1

    Sweden 5 1 0 0 6

    Switzerland 17 0 0 17 0

    Turkey 4 1 1 0 4

    Ukrane 4 0 0 0 4

    UK 56 2 0 58 0

    USA 152 34 110 64 12

    Total 1187 241 162 794 473

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    3

    Geographical variation and phylogeography of an aposematic moth

    https://www.researchgate.net/publication/24036324_Jost_L_GST_and_its_relatives_do_not_measure_differentiation_Mol_Ecol_17_4015-4026?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/24257674_DnaSP_v5_A_Software_for_Comprehensive_Analysis_of_DNA_Polymorphism_Data?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/227464332_Direction_and_strength_of_selection_by_predators_for_the_color_of_the_aposematic_wood_tiger_moth?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/223995688_Peakall_ROD_Smouse_PE_GENALEX_6_genetic_analysis_in_Excel_Population_genetic_software_for_teaching_and_research_Mol_Ecol_Notes_6_288-295?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==https://www.researchgate.net/publication/12483622_STRUCTURE_version_233_inference_of_population_structure_using_multilocus_genotype_data?el=1_x_8&enrichId=rgreq-7779eccf-9f53-4ea6-aee0-6afb3e55ac67&enrichSource=Y292ZXJQYWdlOzI3NTA1NzIzNDtBUzoyMjAxNDA3NTE4NTU2MjJAMTQyOTQ5NzA1NDAyMQ==

  • Species distribution modelling

    Combining climate-based distribution modelling with phylo-

    geographical data allows for development of spatially explicit

    hypotheses of past distribution patterns (Waltari et al., 2007).

    To perform the climatic species distribution model (SDM) of

    P. plantaginis we used Maxent 3.3.3 (Phillips et al., 2006).

    Climate data for the present and for the Last Glacial Maxi-

    mum (LGM) were downloaded from the WorldClim 1.4 data-

    base (Hijmans et al., 2005), using a grid resolution of 2.5

    arcmin. Simulations used to create the climate data for the

    LGM used the Community Climate System Model (CCSM).

    Modelling was replicated 20 times to explore the reliability

    and robustness of the results using 75% of the 211 locality

    records for model training and 25% for model testing (see

    Appendix S1g). Only collection sites with precise locality data

    were used, although such analyses are somewhat robust to

    imprecise locality data (Graham et al., 2008).

    RESULTS

    Hindwing colour distribution

    Hindwing colour varied across the distribution of Parasemia

    plantaginis for both males and females (Fig. 1). Males with

    red hindwings were only found in the Caucasus area. Much

    of the western Palaearctic areas appeared to have a high

    proportion of yellow males, and white males became increas-

    ingly predominant east of Europe, except for interior China.

    We found males with completely melanized hindwings in

    central Siberia and also in the Nearctic.

    Female hindwing colour also exhibited geographical varia-

    tion, but not concordant with male variation. Females with

    deep-red hindwings were most common in the Caucasus.

    However, unlike males, deep-red females can be found in

    much of the western Palaearctic. Eastwards from the Cauca-

    sus, female coloration became predominantly yellow across

    the rest of the Palaearctic and Nearctic. Females with highly

    melanized hindwings could be found occasionally in the

    Nearctic (Fig. 1).

    Phylogenetic relationships and haplotype structure

    Our phylogeny from the COI gene fragment showed shallow

    mitogenetic (mtDNA) structure across most of the Holarctic

    distribution (Fig. 2). Samples from the Caucasus region

    formed a highly supported group (1.0 posterior probability,

    PP). However, one Georgian individual shared a haplotype

    and clustered with moths from central and eastern Russia.

    Despite only moderate haplotype diversity, nucleotide diver-

    sity was greatest in the Caucasus group (Table 2).

    The Bayesian analysis identified two groups in the area

    extending from Europe across to eastern Russia. Two Czech

    moths from a mountainous area formed the first well-sup-

    (a)

    (b)

    (c)

    Figure 1 Hindwing coloration of males andfemales of Parasemia plantaginis across theirgeographical distribution. Colours in circles

    of (a) represent polymorphic male hindwingcolour and (b) represent female hindwing

    colour that varies continuously. Circles in

    (c) enclose sampling sites pooled togetherfor each pie chart based on genetic

    populations (letters A, G, F, E, K, O, P, Q,S, T, U, V, W, X), samples that spatially

    cluster without genetic basis for grouping toraise the sample size above 10 (B, H, I, J,

    M, L, N, R, Y, Z), and a large geneticpopulation split to show greater detail in an

    area adjacent to the Caucasus region (C, D).

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    4

    R. H. Hegna et al.

  • ported group (0.96 PP). The second group, though not sta-

    tistically supported, consisted of samples from central Russia

    to the Kamchatka peninsula on the east coast (0.94 PP).

    However, several specimens within this region had haplo-

    types that grouped among European samples.

    Japanese specimens formed three additional groups rang-

    ing from no support to strong support (0.54–0.99 PP). Oneindividual from the Hokkaido population grouped with a

    population from Honshu and one specimen from Honshu

    shared a haplotype with the European group (Fig. 2).

    North American specimens grouped together, but did not

    form a statistically supported group (0.54 PP, Fig. 2).

    Despite this, no haplotypes were shared with Palaearctic

    moths, or vice versa (Figs 3 & 4). Haplotype diversity within

    the North American group was the highest compared to

    other groups of samples in the analysis (Table 2).

    Our mtDNA haplotype network paralleled results from the

    phylogenetic analysis and also suggested surprisingly little

    differentiation across the Holarctic (Figs 3 & 4). The most

    differentiated haplotypes belonged to individuals from the

    Caucasus. However, one moth from the Caucasus had a hap-

    lotype in common with specimens from Siberia. With the

    exception of a single Japanese and Georgian specimen, all

    other haplotypes were broadly clustered in a geographical

    sense (i.e. Caucasus, Europe–Russia–Japan, and North Amer-ica). No haplotypes were shared between Palaearctic and

    Nearctic specimens.

    Coalescent theory predicts older haplotypes will dominate

    the interior of a network, occur with high frequency, and

    encompass a greater diversity of individuals (Posada & Cran-

    dall, 2001), while younger descendant haplotypes surround

    them and occur at lower frequencies (Crandall & Templeton,

    1993). Consistent with these predictions, our haplotype net-

    work shows one central high-frequency haplotype shared

    across European, Russian, and one Japanese specimen.

    Central haplotypes of relative high frequency also appeared

    in the Caucasus, Russia and North America. Japanese

    moths had several unique haplotypes in addition to the one

    Table 2 Number of samples (n), Tajima’s D and significance level, Fu’s FS value and significance, nucleotide diversity (p), haplotypediversity (h), number of haplotypes (N), and average nucleotide differences within groups (K) calculated from the COI gene fragmentfrom samples of Parasemia plantaginis.

    Group n Tajima’s D P-value FS P-value p h N (K)

    Europe 229 �2.2683 < 0.001 �23.819 < 0.001 0.00093 0.45 20 0.584Caucasus 37 �1.5919 0.033 �0.661 0.379 0.00268 0.428 7 1.686Russia 52 �1.4011 0.037 �2.650 0.053 0.00094 0.501 6 0.591Japan HOK 26 �1.3463 0.052 �1.240 0.130 0.00083 0.348 4 0.520Japan ABO 30 1.0115 0.835 0.605 0.676 0.00172 0.634 4 1.083

    Japan SAJ 23 �1.1533 0.168 0.905 0.693 0.0013 0.53 3 0.822North America 178 �1.3234 0.067 �9.204 0.002 0.00153 0.673 15 0.966Dalton 12 �1.1405 0.133 �0.476 0.333 0.00026 0.167 2 0.167

    Figure 2 Phylogeny showing relationshipsamong Parasemia plantaginis specimensfrom different regions including all 56

    haplotypes of the COI gene fragment.Numbers at nodes are posterior

    probabilities larger than 0.5; values 0.95 andabove are well supported. Red branches

    denote samples from the Caucasus region,black branches are European and Russian

    samples, green branches denote Japanese,and blue branches represent North

    American specimens. Two underlinedspecimens are the haplotypes from the

    Caucasus and Japan that did not group withother haplotypes from those areas. The scale

    bar unit is substitutions per site.

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    5

    Geographical variation and phylogeography of an aposematic moth

  • individual that shared the most common European haplo-

    type. The same patterns in global haplotype structure are evi-

    dent when haplotype frequencies are mapped across

    geographical space (Fig. 3). North American specimens

    exhibited two predominant haplotypes. One haplotype was

    common above the southern Canadian–American border anda second haplotype was predominant below. However, we

    also found each of these two haplotypes to be dominant in

    two different portions of Alaska. Haplotypes were shared by

    males irrespective of hindwing colours (Fig. 4). This lack of

    Figure 3 COI gene fragment haplotypefrequencies of Parasemia plantaginis across

    the entire geographical range of the species.Each colour represents a separate haplotype.

    (a)

    (b)

    Figure 4 Haplotype networks showing eachof the 56 COI haplotypes of Parasemiaplantaginis represented as circles sized in

    proportion to the number of specimenshaving each haplotype. The first network (a)

    shows the geographical origin of sampleswith each haplotype. The second network

    (b) illustrates the proportion of males ofeach haplotype with the four hindwing

    colours observed across their distribution.

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    6

    R. H. Hegna et al.

  • association between hindwing colour and COI haplotypes

    found only in particular geographical regions suggests that

    coloration did not differentiate under a scenario of

    geographical isolation.

    Population structure from microsatellite markers

    Violations of Hardy–Weinberg equilibrium (HWE) or link-age equilibrium were not found across a majority of the

    inferred populations for any loci. The initial analysis of

    HWE and linkage equilibrium with samples grouped by

    country, state, or province also found no violations in any

    loci across a majority of groups. Previous studies indicate the

    influence of null alleles to be negligible at low frequencies

    (< 0.2) (Dakin & Avise, 2004). Our analysis with freenafound that only 23/170 loci by population combinations had

    null alleles with frequencies above 0.2. The influence of null

    alleles on FST values calculated within freena was minimal

    when we compared estimates adjusted for the influence of

    null alleles alongside non-adjusted estimates. The average dif-

    ference was 0.017 and the maximum was 0.06, but none of

    the differences affected the interpretation of FST values (i.e.

    the difference of 0.06 reduced an FST value from 0.48 to

    0.41). Population assignment and delimitation tests, such as

    structure, are also robust to the effects of null alleles, even

    at frequencies as high as 0.9 (Carlsson, 2008). Therefore, our

    ability to infer current populations and examine differences

    among them was not affected.

    Analysis of population structure using structure inferred

    17 populations, with Q > 0.8 across all specimens for theirdetermined population. Populations inferred by structure

    corresponded to some of the strong and weakly supported

    groups in the analysis of the COI gene fragment. These pop-

    ulations included Scotland, central Europe (Europe_C),�Aland, Finland, Turkey, Caucasus (i.e. the northern Caucasus

    Mountains), central Russia (Rus_Central), south-eastern

    Russia (Rus_SE), north-eastern Russia (Rus_NE), the island

    of Hokkaido, Japan (HOK), two populations on Honshu,

    Japan (ABO and SAJ), Fairbanks, Alaska (Alaska_F), south-

    central Alaska (Alaska_K), Canada, Utah and south-eastern

    Utah (Dalton). Populations also tended to correspond to

    large geographical areas, sometimes spanning more than

    200 km wide or more (see Appendix S2c). For instance,

    individuals across Finland grouped together in a population

    spanning 800 km. Specimens from across central Europe

    (Switzerland, Italy, Austria, Czech Republic, Poland, Roma-

    nia, Macedonia and Bulgaria) also formed a single popula-

    tion. Specimens from Estonia indicated a heavily admixed

    ancestry between central Europe and Finland populations

    and did not form their own unique population.

    In contrast to our COI results, analysis of relationships

    among populations using Nei’s genetic distance (DA) in pop-

    tree2 showed a high degree of structure and support for

    many groups only weakly supported in the COI tree (Fig. 5).

    However, none of the structure among populations

    contradicted broad sample groups inferred using the COI

    gene fragment (see Appendix S2d).

    Our analysis of molecular variance attributed 61% of

    molecular variance to within-populations (P < 0.001) and10% to among populations (P < 0.001; Table 3). Among-region variance comprised 29% of molecular variance

    Figure 5 Neighbour-joining tree representing differentiationamong the 17 populations of Parasemia plantaginis, which wederived from analysis of microsatellite markers in structure,

    constructed from genetic distances between populations (DA).Numbers at nodes indicate bootstrap values (1000 bootstrap

    replicates); values above 0.7 indicate strong support for groups.Branches are colour-coded to represent regions as in Fig. 2, with

    Caucasus populations in red, European and Russianpopulations in black, Japanese in green, and North American in

    blue.

    Table 3 Results of a hierarchical analysis of molecular variance (AMOVA) showing the division of genetic variance from themicrosatellite marker data for Parasemia plantaginis across its Holarctic distribution. Regions and populations are defined in the maintext.

    Source d.f. SS MS Est. var. % P-value

    Among regions 5 1072.404 214.481 1.099 29% 0.001

    Among Pops within regions 11 275.938 25.085 0.375 10% 0.001

    Within Pops 1071 2457.633 2.295 2.295 61% 0.001

    Total 1087 3805.974 3.769 1

    d.f., degrees of freedom; SS, sum of squares; MS, mean squares; Est. var., estimated variance.

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    7

    Geographical variation and phylogeography of an aposematic moth

  • (P < 0.001). Regions included Europe, Caucasus, Russia,Japan, Alaska and the southern North American populations

    grouped as regions. All FST values calculated with freena to

    account for null alleles were well supported, having confi-

    dence intervals that did not include zero after 5000 bootstrap

    replicates (Table 4). Our pairwise population FST values indi-

    cate a high degree of differentiation between populations

    separated by 10,000 km (0.48–0.77 when comparing NorthAmerican, Japanese and Caucasus populations), but also sug-

    gest that many neighbouring populations are more moder-

    ately differentiated. Specimens from Caucasus and Turkish

    populations were more strongly differentiated compared with�Aland and Scottish populations (FST = 0.15–0.25) than whencompared with central European or Finland populations

    (FST = 0.12–0.19). While significant, the degree of differenti-ation indicated between the Scottish and central European

    populations was low to moderate (FST = 0.027–0.169).Among Europe and central and eastern Russia, FST values

    indicated moderate differentiation (0.086–0.164). BetweenAlaska and eastern Russia differentiation was high (0.327–0.479). Estimates of DEST showed similar patterns compared

    to FST estimates, except a slightly greater degree of differenti-

    ation among European populations and somewhat less dif-

    ferentiation among North American populations (see

    Appendix S2b).

    Our analysis of gene flow using BayesAss found that

    migration between most population pairs was low enough

    that estimates of recent migration could not be detected

    above background noise. The BayesAss results were consis-

    tent with the populations inferred from structure. Recent

    migration was only detected above background noise

    between two populations in Japan (from ABO to SAJ:

    m = 0.0916, CI95 = 0.043–0.140), and both populations inthe Caucasus region (from Caucasus to Turkey: m = 0.20,CI95 = 0.106–0.290). No evidence of recent migration existedbetween any of the Japanese populations and either south-

    eastern or north-eastern Russia (see Appendix S2e).

    Demographic history and distribution modelling

    Within the Palaearctic, values of Fu’s FS statistic were both

    significant and negative for Europe and Russia, indicative of

    population expansion, but values for other locations were

    not significant (Table 2). In North America, Fu’s FS statistic

    was significant and indicated a negative deviation, except for

    Dalton in southern Utah. Tajima’s D statistic was signifi-

    cantly negative for Europe and the Caucasus, which can sug-

    gest demographic expansion or selection, but was non-

    significantly negative for Hokkaido and North America

    (Table 2).

    The species distribution model suggests that movement

    across the Palaearctic region during the LGM was possible to

    some extent (Fig. 6). The Caucasus region was more con-

    nected to Europe in the past by climatically suitable areas,

    compared to its more isolated present condition. The poten-

    tial for connectivity across large geographical areas appearsTable

    4DifferentiationbetweenParasem

    iaplantaginispopulationsestimated

    bypairw

    iseFSTvalues

    (Weir,1996)adjusted

    fornullallelesusingtheprogram

    freena(below

    thediagonal).

    AllFSTvalues

    had

    confidence

    intervalsnotincludingzero

    afterbootstrapping,

    indicatingthey

    werewellsupported.Genetic

    distance

    betweenpopulations(D

    A)isgivenabove

    thediagonal.

    Population

    Scotland

    Europe_C

    � Aland

    Finland

    Turkey

    Caucasus

    Rus_Central

    Rus_SE

    Rus_NE

    Jap_ABO

    Jap_HOK

    Jap_SA

    JAlaska_F

    Alaska_K

    Canada

    Utah

    Utah_D

    Scotland

    00.250

    0.570

    0.379

    0.538

    0.543

    0.426

    0.460

    0.397

    0.522

    0.495

    0.547

    0.614

    0.631

    0.544

    0.563

    0.565

    Europe_C

    0.027

    00.429

    0.207

    0.496

    0.478

    0.330

    0.429

    0.355

    0.466

    0.424

    0.476

    0.518

    0.546

    0.477

    0.528

    0.524

    � Aland

    0.169

    0.143

    00.291

    0.733

    0.690

    0.647

    0.656

    0.662

    0.686

    0.696

    0.650

    0.747

    0.762

    0.702

    0.767

    0.756

    Finland

    0.068

    0.042

    0.102

    00.551

    0.531

    0.402

    0.407

    0.406

    0.505

    0.445

    0.502

    0.575

    0.629

    0.530

    0.583

    0.570

    Turkey

    0.151

    0.116

    0.211

    0.122

    00.338

    0.467

    0.521

    0.461

    0.629

    0.543

    0.622

    0.650

    0.702

    0.655

    0.648

    0.655

    Caucasus

    0.220

    0.186

    0.252

    0.177

    0.106

    00.549

    0.609

    0.540

    0.724

    0.650

    0.693

    0.714

    0.700

    0.662

    0.690

    0.703

    Rus_Central

    0.120

    0.087

    0.223

    0.109

    0.165

    0.259

    00.264

    0.290

    0.466

    0.311

    0.481

    0.509

    0.530

    0.507

    0.614

    0.584

    Rus_SE

    0.175

    0.150

    0.240

    0.135

    0.242

    0.337

    0.126

    00.303

    0.444

    0.340

    0.459

    0.443

    0.460

    0.418

    0.489

    0.458

    Rus_NE

    0.179

    0.135

    0.259

    0.153

    0.261

    0.332

    0.141

    0.164

    00.304

    0.244

    0.294

    0.511

    0.465

    0.465

    0.481

    0.447

    Jap_ABO

    0.393

    0.307

    0.409

    0.302

    0.562

    0.532

    0.394

    0.373

    0.341

    00.230

    0.104

    0.665

    0.675

    0.662

    0.684

    0.706

    Jap_HOK

    0.339

    0.251

    0.375

    0.247

    0.494

    0.475

    0.288

    0.351

    0.274

    0.262

    00.209

    0.621

    0.608

    0.569

    0.581

    0.611

    Jap_SA

    J0.457

    0.345

    0.443

    0.336

    0.676

    0.581

    0.470

    0.485

    0.400

    0.264

    0.280

    00.676

    0.670

    0.666

    0.702

    0.707

    Alaska_F

    0.418

    0.329

    0.417

    0.332

    0.525

    0.528

    0.446

    0.371

    0.479

    0.620

    0.595

    0.670

    00.073

    0.102

    0.195

    0.174

    Alaska_K

    0.361

    0.280

    0.361

    0.292

    0.487

    0.480

    0.390

    0.327

    0.397

    0.586

    0.560

    0.641

    0.084

    00.135

    0.241

    0.189

    Canada

    0.410

    0.325

    0.445

    0.333

    0.563

    0.559

    0.453

    0.355

    0.471

    0.627

    0.596

    0.672

    0.135

    0.159

    00.121

    0.139

    Utah

    0.491

    0.374

    0.489

    0.380

    0.711

    0.621

    0.558

    0.506

    0.610

    0.711

    0.687

    0.765

    0.333

    0.353

    0.207

    00.146

    Utah_D

    0.445

    0.341

    0.444

    0.341

    0.702

    0.582

    0.514

    0.463

    0.576

    0.712

    0.675

    0.772

    0.268

    0.281

    0.175

    0.418

    0

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    8

    R. H. Hegna et al.

  • to remain high under current conditions except for the

    Caucasus. The average test area under the curve (AUC) of a

    receiver operating characteristic (ROC) across all replicate

    runs was 0.925 (SD = 0.014), which indicated the model out-put was of good quality. The parameters with the highest

    explanatory power included annual mean temperature

    (43.6%), the maximum temperature of the warmest period

    (18.4%), minimum temperature of the coolest period (7%),

    isothermality (6.4%), and annual precipitation (4.4%).

    DISCUSSION

    The wood tiger moth (Parasemia plantaginis) has only two

    main lineages and surprisingly shallow levels of mitogenetic

    structure, given its roughly 17,000 km geographical distribu-

    tion. The first lineage is composed entirely of specimens

    from the Caucasus region and Middle East. The second line-

    age includes specimens from across Europe, Russia, Japan

    and North America, but one individual from the Caucasus

    also has a haplotype in common with this group. Therefore,

    genetic divergence does not completely follow expectations

    based on geographical location, and suggests historical gene

    flow. Hindwing warning colour differences are most strongly

    linked to the split between these two lineages, but such dif-

    ferences are also not exclusive. Males with reddish hindwings

    that vary in the degree of redness occur only in the Caucasus

    region, but yellow males can also be found there on occasion

    (Fig. 1). Female hindwing warning colour is also reddish in

    the Caucasus, but can be red elsewhere in Europe and is pre-

    dominantly yellow across the rest of the distribution east of

    the Ural Mountains. A change in female hindwing colour to

    be predominantly yellow does not appear to correspond with

    any clear genetic differentiation. Therefore, while phenotypic

    differentiation coincides with the greatest mitogenetic diver-

    gence across the distribution in the Caucasus, hindwing col-

    our differences do not appear to be completely exclusive

    between the two groups. Japanese specimens also formed

    their own groups and a shift to predominantly white

    hindwings in males appears to have occurred there. However,

    we observed one yellow individual in Japan and their exis-

    tence is also reported in previous literature (Okano & Katay-

    ama, 1976; Kishida, 2011). Female coloration in Japan is

    similar compared to neighbouring areas (Fig. 1). Therefore,

    genetic structure across the distribution of P. plantaginis

    appears to be somewhat imperfectly matched to the two

    shifts in male coloration in the Caucasus and Japan and not

    associated with female hindwing colour changes. This sug-

    gests that divergence in male and female coloration could be

    the result of selection rather than solely a product of drift

    and long term isolation.

    Shallow mitogenetic structure in lepidopterans across large

    species ranges similar to P. plantaginis has only been

    reported for two other species: the Apollo butterfly, Parnas-

    sius phoebus, which occurs across the Holarctic (Todisco

    et al., 2012), and the pea blue butterfly, Lampides boeticus,

    which occurs between Africa and Australia (Lohman et al.,

    2008). A greater degree of mitogenetic structure has been

    reported for the butterfly Hesperia comma, which is also dis-

    tributed across the Holarctic (Forister et al., 2004). On a

    smaller geographical scale, some studies of other temperate

    and northern lepidopterans have found greater degrees of

    genetic structure and phylogeographical signal (Wahlberg &

    Saccheri, 2007; Gratton et al., 2008; Von Reumont et al.,

    2012).

    Climate oscillations during the Pleistocene that caused

    species’ distributions to contract southwards before the recol-

    onization of northern areas are commonly thought to be a

    predominant driver of mitogenetic structure (Ibrahim et al.,

    1996; Schmitt, 2007; Provan & Bennett, 2008). The influence

    of glacial refugia has been described for many lepidopterans

    (Schmitt, 2007; Kerdelhue et al., 2009). One way of inter-

    preting the deeper split between most of the P. plantaginis

    Caucasus haplotypes and those from the rest of the distribu-

    tion is that the Caucasus served as a glacial refugium, similar

    to what has been hypothesized for other species (Schmitt,

    2007; Kerdelhue et al., 2009). However, our species

    (a)

    (b)

    Figure 6 Species distribution models.Modelled geographical distribution for

    Parasemia plantaginis based on (a) climateunder present conditions showing localities

    used in the model and (b) the modelprojected onto past conditions during the

    Last Glacial Maximum (LGM). Layersprojecting the glacial distribution during the

    LGM in ArcGIS were obtained from Ray &Adams (2001). Areas coloured grey-black

    represent increasing probabilities of climatefavouring the presence of P. plantaginis

    above 10%.

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    9

    Geographical variation and phylogeography of an aposematic moth

  • distribution model (SDM) suggests that suitable climate for

    P. plantaginis existed between southern Europe and the

    Caucasus during the LGM, rather than a disconnected and

    fragmented set of small refugia. Additionally, across Europe

    and Russia we did not find any Caucasus haplotypes, unlike

    patterns of genetic structure in organisms that migrated

    northwards from southern refugia following glacial retreat in

    Europe (Schmitt, 2007; Provan & Bennett, 2008). Less evi-

    dence of population expansion in Russia and in Japan also

    suggests demographic stability during the LGM. Our SDM

    for present conditions, together with molecular data, suggests

    that one explanation for the mitogenetic differentiation

    between the Caucasus and other areas is isolation after post-

    glacial climate change. Exactly when connectivity was lost

    and to what extent gene flow may occur across patches of

    habitable areas in Turkey is unknown. Recent work is reveal-

    ing additional lepidopteran species that show mitogenetic

    structuring not congruent with classic refugia hypotheses

    (Schmitt et al., 2003; Schmitt, 2007; Todisco et al., 2012),

    along with other invertebrates (Crease et al., 2012; Porretta

    et al., 2013). Such patterns appear to be linked to greater

    dispersal abilities of the species in question (Wahlberg &

    Saccheri, 2007). Climate isolations during the Pleistocene

    may have driven selection for greater dispersal abilities in

    northerly taxa (Dynesius & Jansson, 2000), which might help

    explain shallow mitogenetic structure in some Holarctic

    species. Greater potential for past population connectivity

    was also thought to explain the lack of deeply diverged

    clades in the Apollo butterfly across its Holarctic distribution

    (Todisco et al., 2012).

    The deep split between most Caucasus haplotypes and

    other locations, combined with the geographically wide-

    spread predominant European haplotype, suggests that the

    origin of P. plantaginis is based in either south-eastern

    Europe or the Caucasus. Our finding supports previous

    hypotheses suggesting that P. plantaginis has a Eurasian

    origin (Dubatolov, 2008). The question of whether the

    originating populations of P. plantaginis were within the

    Caucasus, or in south-eastern Europe poses fundamental

    questions concerning the evolution of polymorphism and

    sexual dimorphism. Within the European–North Americangroup, males have polymorphic hindwings (yellow or

    white) and females can vary gradually between yellow and

    dark red. However, within the Caucasus group both males

    and females are reddish to varying degrees and very few

    males are yellow. It is interesting that P. plantaginis

    became highly sexually dimorphic over a large portion of

    its distribution because sexual dimorphism does not appear

    in other closely related members of the subfamily Arctiinae

    (Dubatolov, 2008). Given that sexual monomorphism

    appears to be ancestral in arctiines (using phylogenetic

    relationships described in Dubatolov, 2008), we hypothesize

    that the Caucasus population and the red coloration of P.

    plantaginis represents the ancestral form. A more thorough

    understanding of phylogenetic relationships within the Arc-

    tiinae is needed before more detailed inference methods

    can be applied to explore hindwing colour evolution with

    a high degree of confidence.

    Suitable climatic conditions still exist across much of the

    distribution of P. plantaginis according to our SDM and

    should allow for gene flow at some level. This differs

    from the distribution model for the Holarctic-occurring

    Parnassius phoebus (Todisco et al., 2012), which appears to

    be more dependent on cold climate for dispersal than

    P. plantaginis. Despite the possibility for gene flow across

    large areas of suitable climate, we found that recent gene

    flow has not occurred between most populations. High

    levels of current population structure suggest that migration

    among many populations may be rare, but not impossible

    because we detected recent migration between two Japanese

    populations despite the high degree of population differen-

    tiation (FST = 0.26–0.28). Evidence of recent migration inJapan and higher rates of migration into Estonia from both

    central Europe and Finland (but not between central

    Europe and Finland) suggests that some form of selection

    by predators, mates or environment may limit gene

    flow between some populations. The large spatial area

    of genetic populations probably makes intra-population

    connectivity an important influence of male hindwing

    colour polymorphism and female variation. Detailed future

    analyses will be important to understanding how intra-

    population dynamics may influence the hindwing warning

    signal colour.

    What defines a species is not always clear (Mallet, 2008;

    Petit & Excoffier, 2009; Lumley & Sperling, 2011). Our

    results show that populations of P. plantaginis at opposite

    ends of the distribution are highly differentiated, but many

    adjacent populations exhibit more moderated levels of popu-

    lation structure. For instance, Caucasus populations are dif-

    ferentiated the most from Japanese and North American

    populations, but are less differentiated from neighbouring

    European populations (Table 4). Regardless of the level of

    genetic divergence between Caucasus and other populations,

    individuals from the Caucasus are the most diverged in

    hindwing colour. Prezygotic isolation is often held to be a

    critical component in ultimately determining whether intra-

    specific divergence and speciation will take place (Panhuis

    et al., 2001). We know from previous work that the hind-

    wing coloration of males can influence mating success in the

    polymorphic Finnish population (Nokelainen et al., 2012),

    but this has not been tested for the red Caucasus and nearby

    non-red populations in the wild. Lepidopteran pheromones

    can also provide clues concerning species delimitation (Lum-

    ley & Sperling, 2011). During fieldwork in Japan, Canada,

    Alaska, and recently in Caucasus, we found that pheromones

    of Finnish females consistently attracted only P. plantaginis

    males (R.H.H., J.M. & Atsushi Honma, pers. obs.). Further-

    more, we observed no obvious signs of mating hesitancy in

    laboratory crosses between specimens from distant popula-

    tions (n = 59 pairs between Finnish and Japanese individu-als, n = 20 pairs between Caucasus and Finnish individuals;J.M. & Bibiana Rojas, unpublished data). Detailed studies of

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    10

    R. H. Hegna et al.

  • sexual selection in laboratory and field conditions, along with

    better sampling in Turkey and the lower Caucasus range, will

    be needed to fully understand the influence of sexual selec-

    tion on population divergence.

    Overall, our study suggests that recent climatic isolation,

    rather than past isolation during the LGM, is an ongoing

    major influence in the genetic structure of P. plantaginis at

    continental scales, along with sexual selection (Nokelainen

    et al., 2012) and predation (Nokelainen et al., 2012; Hegna

    et al., 2013; Hegna & Mappes, 2014) at other spatial scales.

    Indications of more recent genetic differentiation, probably

    due to the combination of these factors, raises the possibility

    that incipient speciation is taking place in portions of the

    distribution, such as the Caucasus. The apparent shift from

    sexual monomorphism to distinct sexual dimorphism in col-

    oration outside the Caucasus region raises the possibility that

    mate choice, in addition to isolation, may have contributed

    to the divergence of male coloration. Prezygotic isolation

    may be a factor maintaining phenotypic and genetic

    differences in the Caucasus if immigration from Europe does

    happen periodically, but only direct experimentation will be

    able to confirm mate choice. It is also possible that predator-

    driven selection on the hindwing warning signal is more

    intensive in the Caucasus, favouring warning signal mono-

    morphism, as is expected for aposematic species (Joron &

    Mallet, 1998). Future work with P. plantaginis promises to

    contribute to our understanding of the evolution of colour

    polymorphism, sexual dimorphism, and warning coloration.

    ACKNOWLEDGEMENTS

    We thank the following people who assisted in the collection

    of moths on such a large scale: Ossi Nokelainen, T€onis Ta-

    sane, Lauri Mikonranta, Kari Kulmala, Otmar Wanninger,

    Felix Sperling, Jordi Maggi, Reza Zahiri, Kyle Johnson, Jona-

    than Hegna, Chuck Harp, Evgenij Derzhinski, Sergey Sinev,

    Alexej Matov, Venera Tyukmaeva, Ted Pike, Jostein Engdal,

    Julien Delisle, Vladamir Gusarov, Leif Aarvik, and Robert

    Mower. Sari Viinikainen, Carol Gilsenan, and Roghaieh

    Ashrafi assisted in lab work. We thank the following muse-

    ums for loaning collections or allowing us to visit: the Zoo-

    logical Museum of the Zoological Institute of the Russian

    Academy of Sciences, University of Oslo, National Royal

    Museum of Sweden, University of Helsinki, the American

    Museum of Natural History, University of Alberta, Brigham

    Young University, University of Wisconsin-Madison, and the

    Finnish Museum of Natural History. We also thank Emily

    Knott, Niklas Wahlberg, Kyle Summers, Mathieu Joron, Tho-

    mas Jones, and Swanne Gordon who improved the quality of

    this manuscript with their comments. Permits included: the

    Swiss National Park (permit no. CH-4168), Hohe Tauen

    National Park (permit no. 21303-96/8/11-2010, permit no.

    SP3-NS-1545/2010), Waterton National Park (WL-2009-

    2915), US Fish and Wildlife Service (KN-11-001), and Alaska

    Department of Fish and Game (FH11-III-002-SA). Finnish

    Centre of Excellence in Biological Interactions 2012–2017

    (no. 252411) and BIOINT graduate school provided funding

    (R.H.H.).

    REFERENCES

    Carlsson, J. (2008) Effects of microsatellite null alleles on

    assignment testing. Journal of Heredity, 99, 616–623.Chapuis, M.P. & Estoup, A. (2007) Microsatellite null alleles

    and estimation of population differentiation. Molecular

    Biology and Evolution, 24, 621–631.Chouteau, M. & Angers, B. (2012) Wright’s shifting balance

    theory and the diversification of aposematic signals. PLoS

    ONE, 7, e34028.

    Clement, M., Posada, D. & Crandall, K.A. (2000) TCS: a

    computer program to estimate gene genealogies. Molecular

    Ecology, 9, 1657–1659.Coolidge, K.R. (1910) Western Lepidoptera.—III.—Notes on

    Leptarctia californiae. Psyche, 17, 166–168.Crandall, K.A. & Templeton, A.R. (1993) Empirical tests of

    some predictions from coalescent theory with applications

    to intraspecific phylogeny reconstruction. Genetics, 134,

    959–69.Crease, T.J., Omilian, A.R., Costanzo, K.S. & Taylor, D.J.

    (2012) Transcontinental phylogeography of the Daphnia

    pulex species complex. PLoS ONE, 7, e46620.

    Dakin, E.E. & Avise, J.C. (2004) Microsatellite null alleles in

    parentage analysis. Heredity, 93, 504–509.Darriba, D., Taboada, G.L., Doallo, R. & Posada, D. (2012)

    jModelTest 2: more models, new heuristics and parallel

    computing. Nature Methods, 9, 772–772.De Freina, J.J. (1993) Contribution to the knowledge of the

    Balkan and West Asian populations of Parasemia plantaginis

    (Linnaeus, 1758) with the description of a new subspecies

    from the Kurdish Zagros-range: carbonelli n. ssp. (Lepidop-

    tera, Arctiidae). Linneana Belgica, 14, 155–164.Dubatolov, V.V. (2008) Construction of the phylogenetic

    model for the genera of the tribe Arctiini (Lepidoptera,

    Arctiidae) with the SYNAP method. Entomological Review,

    88, 833–837.Dubatolov, V.V. & Zahiri, R. (2005) Tiger-moths of Iran.

    Atalanta, 36, 481–525.Dynesius, M. & Jansson, R. (2000) Evolutionary conse-

    quences of changes in species’ geographical distributions

    driven by Milankovitch climate oscillations. Proceedings of

    the National Academy of Sciences USA, 97, 9115–9120.Endler, J.A. (1977) Geographic variation, speciation and clines.

    (MPB-10). Princeton University Press, Princeton, NJ.

    Forister, M.L., Fordyce, J.A. & Shapiro, A.M. (2004) Geologi-

    cal barriers and restricted gene flow in the holarctic skip-

    per Hesperia comma (Hesperiidae). Molecular Ecology, 13,

    3489–3499.Galarza, J.A., Viinikainen, S.M., Ashrafi, R. & Mappes, J.

    (2010) First microsatellite panel for the Wood Tiger Moth

    (Parasemia plantaginis). Conservation Genetics Resources, 3,

    197–199.

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    11

    Geographical variation and phylogeography of an aposematic moth

  • Galarza, J.A., Nokelainen, O., Ashrafi, R., Hegna, R.H. &

    Mappes, J. (2014) Temporal relationship between genetic

    and warning signal variation in the aposematic wood tiger

    moth (Parasemia plantaginis). Molecular Ecology, 23, 4939–4957.

    Graham, C.H., Elith, J., Hijmans, R.J., Guisan, A., Peterson,

    A.T. & Loiselle, B.A. & The Nceas Predicting Species Dis-

    tributions Working Group (2008) The influence of spatial

    errors in species occurrence data used in distribution

    models. Journal of Applied Ecology, 45, 239–247.Gratton, P., Konopi�nski, M.K. & Sbordoni, V. (2008) Pleisto-

    cene evolutionary history of the Clouded Apollo (Parnas-

    sius mnemosyne): genetic signatures of climate cycles and a

    ‘time-dependent’ mitochondrial substitution rate. Molecu-

    lar Ecology, 17, 4248–4262.H€arlin, C. & H€arlin, M. (2003) Towards a historization of

    aposematism. Evolutionary Ecology, 17, 197–212.Hegna, R.H. & Mappes, J. (2014) Influences of geographic

    differentiation in the forewing warning signal of the

    wood tiger moth in Alaska. Evolutionary Ecology, 28,

    1003–1017.Hegna, R.H., Nokelainen, O., Hegna, J.R. & Mappes, J.

    (2013) To quiver or to shiver: increased melanization ben-

    efits thermoregulation, but reduces warning signal efficacy

    in the wood tiger moth. Proceedings of the Royal Society B:

    Biological Sciences, 280, 20122812.

    Hijmans, R.J., Cameron, S.E., Parra, J.L., Jones, P.G. & Jarvis,

    A. (2005) Very high resolution interpolated climate sur-

    faces for global land areas. International Journal of Clima-

    tology, 25, 1965–1978.Hines, H.M., Counterman, B.A., Papa, R., Albuquerque de

    Moura, P., Cardoso, M.Z., Linares, M., Mallet, J., Reed,

    R.D., Jiggins, C.D., Kronforst, M.R. & McMillan, W.O.

    (2011) Wing patterning gene redefines the mimetic history

    of Heliconius butterflies. Proceedings of the National Acad-

    emy of Sciences USA, 108, 19666–19671.Ibrahim, K.M., Nichols, R.A. & Hewitt, G.M. (1996) Spatial

    patterns of genetic variation generated by different forms

    of dispersal. Heredity, 77, 282–291.Joron, M. & Mallet, J.L.B. (1998) Diversity in mimicry: para-

    dox or paradigm? Trends in Ecology and Evolution, 13,

    461–466.Jost, L.O.U. (2008) GST and its relatives do not measure dif-

    ferentiation. Molecular Ecology, 17, 4015–4026.Kapan, D.D. (2001) Three-butterfly system provides a field

    test of mullerian mimicry. Nature, 409, 338–340.Kerdelhue, C., Zane, L., Simonato, M., Salvato, P., Rousselet,

    J., Roques, A. & Battisti, A. (2009) Quaternary history and

    contemporary patterns in a currently expanding species.

    BMC Evolutionary Biology, 9, 220.

    Kishida, Y. (2011) The standard of moths in Japan II (ed. by

    Y. Kishida), p. 162. Gakken Education Press, Tokyo (in

    Japanese).

    Koc�ak, A.€O. (1989) Notes on the Lepidoptera of ErzurumProvince. Priamus, 5, 73–104.

    Librado, P. & Rozas, J. (2009) DnaSP v5: a software for com-

    prehensive analysis of DNA polymorphism data. Bioinfor-

    matics, 25, 1451–1452.Lindstedt, C., Eager, H., Ihalainen, E., Kahilainen, A., Ste-

    vens, M. & Mappes, J. (2011) Direction and strength of

    selection by predators for the color of the aposematic

    wood tiger moth. Behavioral Ecology, 22, 580–587.Lohman, D.J., Peggie, D., Pierce, N.E. & Meier, R. (2008)

    Phylogeography and genetic diversity of a widespread Old

    World butterfly, Lampides boeticus (Lepidoptera: Lycaeni-

    dae). BMC Evolutionary Biology, 8, 301.

    Lumley, L.M. & Sperling, F.A.H. (2011) Utility of microsatel-

    lites and mitochondrial DNA for species delimitation in

    the spruce budworm (Choristoneura fumiferana) species

    complex (Lepidoptera: Tortricidae). Molecular Phylogenet-

    ics and Evolution, 58, 232–243.Mallet, J. (2008) Hybridization, ecological races and the nat-

    ure of species: empirical evidence for the ease of specia-

    tion. Philosophical Transactions of the Royal Society B:

    Biological Sciences, 363, 2971–2986.Nei, M., Tajima, F. & Tateno, Y. (1983) Accuracy of esti-

    mated phylogenetic trees from molecular data. Journal of

    Molecular Evolution, 19, 153–170.Nokelainen, O., Hegna, R.H., Reudler, J.H., Lindstedt, C. &

    Mappes, J. (2012) Trade-off between warning signal effi-

    cacy and mating success in the wood tiger moth. Proceed-

    ings of the Royal Society B: Biological Sciences, 279, 257–265.

    Nokelainen, O., Lindstedt, C. & Mappes, J. (2013) Environ-

    ment-mediated morph-linked immune and life-history

    responses in the aposematic wood tiger moth. Journal of

    Animal Ecology, 82, 653–662.Okano, M. & Katayama, C. (1976) A revision of the Japanese

    subspecies and forms of Parasemia plantaginis (Linnaeus)

    (Lepidoptera: Arctiidae). Artes Liberales, 18, 41–46.Panhuis, T.M., Butlin, R., Zuk, M. & Tregenza, T. (2001)

    Sexual selection and speciation. Trends in Ecology and Evo-

    lution, 16, 364–371.Peakall, R. & Smouse, P.E. (2006) GENALEX 6: genetic

    analysis in Excel. Population genetic software for teaching

    and research. Molecular Ecology Notes, 6, 288–295.Petit, R.J. & Excoffier, L. (2009) Gene flow and species

    delimitation. Trends in Ecology and Evolution, 24, 386–393.

    Phillips, S.J., Anderson, R.P. & Schapire, R.E. (2006) Maxi-

    mum entropy modeling of species geographic distribu-

    tions. Ecological Modelling, 190, 231–259.Porretta, D., Mastrantonio, V., Mona, S., Epis, S., Montagna,

    M., Sassera, D., Bandi, C. & Urbanelli, S. (2013) The inte-

    gration of multiple independent data reveals an unusual

    response to Pleistocene climatic changes in the hard tick

    Ixodes ricinus. Molecular Ecology, 22, 1666–1682.Posada, D. & Crandall, K.A. (2001) Intraspecific gene geneal-

    ogies: trees grafting into networks. Trends in Ecology and

    Evolution, 16, 37–45.

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    12

    R. H. Hegna et al.

  • Pritchard, J.K., Stephens, M. & Donnelly, P. (2000) Inference

    of population structure using multilocus genotype data.

    Genetics, 155, 945–959.Provan, J. & Bennett, K.D. (2008) Phylogeographic insights

    into cryptic glacial refugia. Trends in Ecology and Evolu-

    tion, 23, 564–571.Rambaut, A. & Drummond, A. (2003) Tracer: a program for

    analysing results from Bayesian MCMC programs such as

    BEAST & MrBayes. University of Edinburgh, UK. Available

    at: http://beast.bio.ed.ac.uk/Tracer.

    Ray, N. & Adams, J.M. (2001) A GIS-based vegetation map

    of the world at the last glacial maximum (25,000–15,000BP). Internet Archaeology, 11, 1–44.

    Ronquist, F. & Huelsenbeck, J.P. (2003) MrBayes 3: Bayesian

    phylogenetic inference under mixed models. Bioinformat-

    ics, 19, 1572–1574.Schemske, D.W., Mittelbach, G.G., Cornell, H.V., Sobel, J.M.

    & Roy, K. (2009) Is there a latitudinal gradient in the

    importance of biotic interactions? Annual Review of Ecol-

    ogy, Evolution, and Systematics, 40, 245–269.Schmitt, T. (2007) Molecular biogeography of Europe: Pleis-

    tocene cycles and postglacial trends. Frontiers in Zoology,

    4, 1–13.Schmitt, T., Giessl, A. & Seitz, A. (2003) Did Polyommatus

    icarus (Lepidoptera: Lycaenidae) have distinct glacial

    refugia in southern Europe? Evidence from population

    genetics. Biological Journal of the Linnean Society, 80,

    529–538.Takezaki, N., Nei, M. & Tamura, K. (2010) POPTREE2: soft-

    ware for constructing population trees from allele fre-

    quency data and computing other population statistics

    with Windows interface. Molecular Biology and Evolution,

    27, 747–752.Todisco, V., Gratton, P., Zakharov, E.V., Wheat, C.W., Sbor-

    doni, V. & Sperling, F.A.H. (2012) Mitochondrial phyloge-

    ography of the Holarctic Parnassius phoebus complex

    supports a recent refugial model for alpine butterflies.

    Journal of Biogeography, 39, 1058–1072.Von Reumont, B.M., Struwe, J.-F., Schwarzer, J. & Misof,

    B. (2012) Phylogeography of the burnet moth Zygaena

    transalpina complex: molecular and morphometric differ-

    entiation suggests glacial refugia in Southern France,

    Western France and micro-refugia within the Alps. Jour-

    nal of Zoological Systematics and Evolutionary Research,

    50, 38–50.

    Wahlberg, N. & Saccheri, I. (2007) The effects of Pleistocene

    glaciations on the phylogeography of Melitaea cinxia (Lep-

    idoptera: Nymphalidae). European Journal of Entomology,

    104, 675.

    Waltari, E., Hijmans, R.J., Peterson, A.T., Ny�ari, �A.S., Per-

    kins, S.L. & Guralnick, R.P. (2007) Locating Pleistocene

    refugia: comparing phylogeographic and ecological niche

    model predictions. PLoS ONE, 2, e563.

    Wang, I.J. & Shaffer, H.B. (2008) Rapid color evolution in

    an aposematic species: a phylogenetic analysis of color var-

    iation in the strikingly polymorphic strawberry poison-

    dart frog. Evolution, 62, 2742–2759.Weir, B. (1996) Genetic data analysis II. Sinauer Associates,

    Sunderland, MA.

    Wilson, G.A. & Rannala, B. (2003) Bayesian inference of

    recent migration rates using multilocus genotypes. Genet-

    ics, 163, 1177–1191.

    SUPPORTING INFORMATION

    Additional Supporting Information may be found in the

    online version of this article:

    Appendix S1 Supplementary methods.

    Appendix S2 Supplementary results.

    BIOSKETCHES

    Robert Hegna is a former PhD student at the University of

    Jyv€askyl€a in Finland. His research interests include the evolu-

    tionary ecology and behaviour of aposematic species.

    Juan Galarza is currently a postdoc at the University of

    Jyv€askyl€a working on the evolution of aposematism, and the

    ecology of insect social behaviour.

    Johanna Mappes is a professor of evolutionary ecology at

    the University of Jyv€askyl€a interested in evolution of animal

    signalling and predator–prey interactions. She leads the Cen-tre of Excellence in Biological Interactions research group.

    Author contributions: J.M., R.H.H. and J.A.G. conceived the

    ideas; R.H.H. collected the data; R.H.H. and J.A.G. analysed

    the data; and R.H.H. led the writing.

    Editor: Luiz Rocha

    Journal of Biogeographyª 2015 John Wiley & Sons Ltd

    13

    Geographical variation and phylogeography of an aposematic moth

    All in-text references underlined in blue are linked to publications on ResearchGate, letting you access and read them immediately.

    http://beast.bio.ed.ac.uk/Tracer