Skip to main content
Advertisement
Browse Subject Areas
?

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

For more information about PLOS Subject Areas, click here.

  • Loading metrics

A Comparative Analysis of Genetic Differentiation across Six Shared Willow Host Species in Leaf- and Bud-Galling Sawflies

  • Sanna A. Leppänen ,

    sanna.a.leppanen@uef.fi

    Affiliation Department of Biology, University of Eastern Finland, P. O. Box 111, FI-80101, Joensuu, Finland

  • Tobias Malm,

    Affiliation Department of Zoology, Swedish Museum of Natural History, P. O. Box 50007, SE-10405, Stockholm, Sweden

  • Kaisa Värri,

    Affiliation Department of Biology, University of Eastern Finland, P. O. Box 111, FI-80101, Joensuu, Finland

  • Tommi Nyman

    Affiliation Department of Biology, University of Eastern Finland, P. O. Box 111, FI-80101, Joensuu, Finland

Abstract

Genetic divergence and speciation in plant-feeding insects could be driven by contrasting selection pressures imposed by different plant species and taxa. While numerous examples of host-associated differentiation (HAD) have been found, the overall importance of HAD in insect diversification remains unclear, as few studies have investigated its frequency in relation to all speciation events. One promising way to infer the prevalence and repeatability of HAD is to estimate genetic differentiation in multiple insect taxa that use the same set of hosts. To this end, we measured and compared variation in mitochondrial COI and nuclear ITS2 sequences in population samples of leaf-galling Pontania and bud-galling Euura sawflies (Hymenoptera: Tenthredinidae) collected from six Salix species in two replicate locations in northern Fennoscandia. We found evidence of frequent HAD in both species complexes, as individuals from the same willow species tended to cluster together on both mitochondrial and nuclear phylogenetic trees. Although few fixed differences among the putative species were found, hierarchical AMOVAs showed that most of the genetic variation in the samples was explained by host species rather than by sampling location. Nevertheless, the levels of HAD measured across specific pairs of host species were not correlated in the two focal galler groups. Hence, our results support the hypothesis of HAD as a central force in herbivore speciation, but also indicate that evolutionary trajectories are only weakly repeatable even in temporally overlapping radiations of related insect taxa.

Introduction

The extreme species richness of plant-feeding insects has puzzled researchers for decades [1][3]. One possible driver of insect herbivore diversification is that shifts onto novel host plants lead to the formation of new species [4][6]. This could occur if chemical, ecological, morphological, and phenological differences among plant taxa act as a source of divergent natural selection [7][9]. In such cases, host-associated differentiation (HAD) in insect populations can lead to the formation of reproductively partially isolated host races [10], [11] that eventually split into fully separated lineages, each of which is associated with a different plant [12], [13].

Genetic studies have revealed numerous cases of HAD among insect populations and species [3], [10], [14]. However, it is still unknown whether HAD constitutes a generally important driver of herbivore diversification, because this would require knowledge on its frequency in relation to all speciation events [15][17]. One way to infer the prevalence of HAD is to estimate genetic divergence in multiple insect taxa that share the same set of host plants: if HAD is common, it should be observable in many of the focal groups [12]. Further, if the buildup of reproductive isolation is predictable, the level of HAD across specific host species and taxa should be roughly equal in independently evolving insect lineages, especially if they are of similar age. Unfortunately, few studies have investigated HAD in more than one herbivore–host taxon pair, meaning that while the literature abounds with convincing case studies, an understanding of the frequency and repeatability of HAD is still lacking [12], [18], [19].

An exceptionally suitable model system for studying parallel HAD is provided by the species-rich network formed by willows (Salicaceae: Salix spp.) and willow-galling sawflies that belong to the nematine subtribe Euurina (Hymenoptera: Tenthredinidae). At approximately 300–500 species, willows constitute one of the most diverse plant genera in the Holarctic region [20][22]. The circa 400 species of Euurina gallers induce leaf folds or rolls, or various types of closed galls on leaves, petioles, buds, or shoots of Salix species (only a handful of leaf-rolling/folding species occur on Populus in the same plant family) [23], [24]. Phylogeny-based studies have revealed that gall-type shifts have been relatively rare during the evolutionary history of the Euurina, while the use of Salix species has been far more labile [25], [26]. Therefore, extant Euurina species represent the end product of multiple, temporally overlapping radiations of lineages representing different gall morphologies. Another consequence is that willows and Euurina gallers occur in multispecies communities, in which each willow species typically is attacked by numerous sawfly species that induce different galls. Based on subtle morphological differences, most Euurina species are presumed to be strictly monophagous on single Salix species [23], [24], but the proposition has rarely been tested using population-genetic methods [27]. As a result, the level of reproductive isolation – and resultant genetic divergence – among the putative species remains largely unstudied.

Our present study focused on leaf- and bud-galling sawflies belonging to the Pontania viminalis-group and the Euura mucronata-group, respectively [23], [24], [28] (Fig. 1). These two taxa often co-occur on sympatric Salix species and, therefore, can be considered independent replicates of differentiation and radiation across the same ‘selective background’ formed by willows. Both groups are composed of univoltine closely related species that in many cases are very difficult to separate on the basis of morphological traits [24], [26], [29]. Within Europe, Kopelke [23], [28] recognized 19 host-specific species in the P. viminalis-group and 16 species in the E. mucronata-group. Kopelke's hypothesis of a strict 1∶1 relationship between willows and various groups of Euurina gallers has, however, been genetically investigated only in northern European Euura bud gallers, in which Nyman [27], using allozyme-electrophoretic data, found evidence for the existence of three or four variably specialized lineages across six willow host species. However, the degree of genetic differentiation and reproductive isolation among the lineages remained unclear, and comparable data on the leaf-galling P. viminalis-group are still lacking.

thumbnail
Figure 1. Examples of galls induced by the focal sawfly groups on willows.

(A) Pontania leaf gall on Salix glauca, and (B) Euura bud gall on the same host species (upper bud galled, lower one uninfected).

https://doi.org/10.1371/journal.pone.0116286.g001

In order to study the repeatability of HAD in shared ecological settings, we collected Pontania and Euura individuals from galls on six willow species that are used by both taxa at two replicate locations in northern Fennoscandia (Abisko in Sweden and Kilpisjärvi in Finland), and then used DNA sequence data from the mitochondrial COI gene and the nuclear ITS2 region to infer their levels of genetic differentiation across host species and geographic locations. Specifically, we wanted to find out: (1) is there evidence for HAD in the two focal species complexes, and (2) is the level of HAD across pairs of willow host species similar in bud- and leaf-galling sawflies?

Methods

Sample collection

Pontania and Euura samples were collected from the surroundings of the Kilpisjärvi research station (69°03′N, 20°48′E) in Finland and the Abisko research station (68°21′N, 18°49′E) in Sweden. Both sites are located c. 250 km north of the Arctic Circle, and the distance between them is about 110 km. The stations are surrounded by very similar habitats, with lowlands dominated by mountain birch (Betula pubescens ssp. czerepanovii (N. I. Orlova) Hämet-Ahti) forests, and a subalpine zone consisting of Salix thickets surrounding the tree- and shrubless higher areas and fjell tops. The focal taxa are very common, and no specific permissions were required for collecting samples.

Galler population samples from specific willow species were collected haphazardly and at intervals of at least 20 meters, in order to avoid sampling of closely related individuals. Euura bud galls were collected in 1998 from six willow species that occur in both areas: Salix lanata L., S. glauca L., S. lapponum L., S. phylicifolia L., S. myrsinifolia Salisb., and S. hastata L. (Table 1). The larvae were reared to adults and sexed, and then used in the allozyme-electrophoretic study by Nyman [27]. The remains of the individuals were stored frozen in 1.5-ml Eppendorf tubes at −80°C until DNA extraction and sequencing. Only Euura males were used in this study: because hymenopteran males are haploid, nuclear sequences could be determined to the haplotype level. Pontania larvae were collected in 1997 and 1998 from the aforementioned six willow species, but also from S. borealis (Fr.) Nasarov and S. reticulata L. In addition, Pontania samples were collected from S. myrsinites L. in Kilpisjärvi, but leaf galls were absent from this host in Abisko. Pontania specimens were stored in 99.5% ethanol at −20°C, but the sex of the larvae could not be determined.

thumbnail
Table 1. Numbers of mitochondrial COI gene and nuclear ITS2 region sequences obtained for leaf-galling Pontania and bud-galling Euura individuals originating from different willow host species and collection sites.

https://doi.org/10.1371/journal.pone.0116286.t001

We aimed at collecting and sequencing ten sawfly individuals per host species from both locations, but for some of the hosts or sites the final sample sizes were lower due to rarity of galls and/or problems in PCR amplification (Table 1). The total number of sequenced individuals was 245.

DNA extraction, PCR amplification, sequencing, and sequence alignment

Total genomic DNA was extracted from the samples using DNeasy Tissue Kits (Qiagen, Valencia, CA). A portion of the mitochondrial Cytochrome oxidase I (COI) gene was amplified and sequenced using methods described earlier [30]. In addition, we sequenced a fragment of the nuclear Internal transcribed spacer 2 (ITS2) region, which was amplified using primers AM1 (5′ TGT GAA CTG CAG GAC ACA TGA 3′) and AM2 (5′ ATG CTT AAA TTT AGG GGG TAG TC 3′) [31]. ITS2 amplification and sequencing was done following the same protocols as for COI [30]. The PCR program for ITS2 consisted of an initial denaturing step at 94°C for 5 min, followed by 35 cycles of 1 min denaturing at 94°C, 1 min annealing at 52°C, and 1 min extension at 72°C. We also designed four new internal primers that were used when needed to confirm the sequences of the PCR products: AMfor180 (5′CGA TCA AAG ACT TGT ACA CG 3′), AMfor250 (5′ GTC GCG TCC CAG TGC TAT CTG 3′), AMrev430 (5′ACG AGG ATA AAT TCG ACT ACC 3′), and AMrev460 (5′ GAC GGT CGG ACG TGT ACG T 3′).

The 238 COI sequences were all 810 bp in length, and were read, edited, and aligned using Sequencher version 4.8 (Gene Codes Corp., Ann Arbor, MI). Thirteen cases of heteroplasmic COI sequences were observed in Pontania and six in Euura (Table 1); we note that since no stop codons were present in the sequences, these instances are more likely to reflect true intra-individual heteroplasmy rather than the presence of nuclear pseudogenes (cf. [32]), so the sequences were retained in our analyses. ITS2 sequences of 222 individuals were aligned by implementing ClustalW [33] in Mesquite version 2.74 [34], followed by minor manual correction; the final alignment is 611 bp long. All sequences have been deposited in GenBank under accession numbers KJ155839–KJ156298.

Phylogeny reconstruction

The COI, ITS2, and combined COI+ITS2 datasets were analyzed in a Bayesian phylogenetic framework in MrBayes version 3.2.2 [35] on the CIPRES Science Gateway [36]. Specimens lacking data for either one of the genes were excluded from the combined-data analysis, making the final number of individuals in the COI+ITS2 dataset 215. Because the ITS2 alignment contained gaps, which are treated as unknowns by model-based methods of phylogenetic inference, gaps were recoded to binary presence/absence characters following the method of Simmons and Ochoterena [37] as implemented in FastGap version 1.2 [38]. In all analyses, we used a mixed substitution model with gamma-categorized rate variation for nucleotide sites, while the recoded gaps were treated as a separate partition of variable, gamma-corrected restriction sites. Ratepr, statefreqpr, shape, revmat, and tratio were left as default and unlinked across partitions. Analyses were run over four chains for 10 million generations while sampling trees every 1,000 generations; run convergence was assessed in Tracer v.1.6.0 [39] to ensure that the 25% burn-in cutoffs included all pre-convergence samples. All trees were midpoint-rooted, following results by [25] and [26].

Analyses of genetic differentiation

ARLEQUIN version 3.5 [40] was used to estimate pairwise ΦST distances among population samples collected from different hosts and locations, as well as to conduct hierarchical analyses of molecular variance (AMOVA) [41]. Calculations were performed separately for the mitochondrial and nuclear sequence datasets. The best-fitting substitution model for COI ( = TrN+I = TN93+I) was determined under the corrected Akaike Information Criterion (AICc) in jModelTest version 2.1.3 [42]. Because a proportion of invariant sites cannot be included in analyses in ARLEQUIN, we instead used the TN93 substitution model with a gamma correction for rate heterogeneity (α = 0.011, estimated by jModelTest). Analyses of the ITS2 data were based on numbers of pairwise differences among sequences, and gaps were weighted equally to substitutions. As our main aim was to compare levels of HAD in Euura and Pontania gallers, we included in these analyses only samples originating from the six willow host species that are shared by both groups. Hence, Pontania individuals sampled from S. myrsinites and S. reticulata were excluded, and S. borealis-associated Pontania larvae were united under S. myrsinifolia; S. borealis can be considered a subspecies of the latter (S. myrsinifolia ssp. borealis (Fr.) Hyl.; see [43]), and the population samples collected from these two willows did not differ when samples were pooled across locations (COI: ΦST = −0.033, P = 0.472; ITS2: ΦST = 0.000, P = 1.000). Hierarchical AMOVAs for genetic differentiation across the population samples were performed using two different grouping hierarchies: collection locations within host species, and vice versa. Statistical significance of estimated parameters was evaluated using 10,000 permutations.

Differences in average ΦST between Pontania and Euura were analyzed with paired-samples t-tests in IBM SPSS version 19.0 (IBM SPSS statistics Inc., Chicago, IL, USA). In order to compare ΦST values across host-species pairs, populations sharing hosts in different locations were pooled, and the existence of a correlation between estimates based on the Pontania and Euura datasets was assessed using a Mantel test employing 10,000 randomizations in PC-ORD version 5.0 [44]. Correlations between among-host differentiation in COI and ITS2 within the two genera were tested similarly.

Results

Phylogenetic trees

Midpoint-rooted 50% majority-rule Bayesian trees show for both COI and ITS2 a highly supported split between Pontania and Euura, and long branches leading to each genus (insets in Fig. 2). By contrast, genetic distances within the genera are very short, and many groupings therefore remain unresolved or receive weak support (i.e., posterior probability <90%). Nevertheless, all single-gene trees demonstrate clear non-random clustering according to host plant in both focal species complexes.

thumbnail
Figure 2. Bayesian phylogenetic trees for mitochondrial COI sequences in (A) Pontania and (B) Euura, and for nuclear ITS2 sequences in (C) Pontania and (D) Euura.

Outlines of full midpoint-rooted trees for both genes are shown in the insets, host plants are indicated by individual label colours and symbols, and collection sites by symbol fills (see legend). Numbers above branches are posterior probabilities (%), the suffix ‘2 h’ denotes individuals with heteroplasmic COI or heterozygous ITS2 sequences.

https://doi.org/10.1371/journal.pone.0116286.g002

In the Pontania branch of the COI tree (Fig. 2A), no monophyletic groups are formed exclusively by all specimens collected from a single willow host species. However, many clusters consisting of identical or near-identical sequences are dominated by samples from one host species. The Euura COI branch (Fig. 2B) shows less resolution, but all individuals collected from S. lapponum are grouped within a single clade. Larvae collected from S. phylicifolia and S. myrsinifolia are found in several groups next to the ‘S. lapponum clade,’ but these hosts are also represented in the large, tight clade containing all specimens from S. lanata, S. hastata, and S. glauca.

The Pontania half of the ITS2 tree (Fig. 2C) exhibits somewhat clearer host-based grouping than does the corresponding part of the COI tree. Individuals from S. lapponum and S. phylicifolia form distinct groups, and marked host-based frequency differences are found also in the other clades of the tree, although most specimens collected from S. myrsinifolia, S. borealis, and S. glauca share a single sequence type. Likewise, the Euura branch of the ITS2 tree (Fig. 2D) is more structured than the Euura COI phylogeny. Individuals from S. lapponum form a strongly supported and near-exclusive clade. Notably, individuals from S. lanata and S. glauca occur in many places across the tree, but are often intermixed in the same clades, and a similar pattern is discernible for samples collected from S. myrsinifolia and S. phylicifolia.

The overall structure of the combined-data tree (inset in Fig. 3) closely resembles the results of the separate analyses of the COI and ITS2 datasets. The Pontania half of the tree (Fig. 3A) groups individuals according to host plants clearer than do the single-gene trees. Individuals originating from S. reticulata form a single clade, and individuals from S. phylicifolia, S. lanata, S. hastata, and S. lapponum occur in two separate, yet exclusive groups each. Specimens collected from S. myrsinifolia and S. borealis tend to co-occur in clades, while samples from S. glauca are scattered in many clades across the tree. Within Euura (Fig. 3B), the combined-data tree largely reflects the COI partition of the dataset, yet provides added resolution as compared to the COI tree. All S. lapponum-associated individuals are included in a single monophyletic group. Individuals collected from S. lanata and S. glauca are intermixed in several shared groups. S. hastata-associated specimens occur in several groups, but nearly all of them have very similar overall sequences. Samples from S. phylicifolia and S. myrsinifolia are intermixed and found in two relatively distant branches on the tree, which seems to be driven by their occurrence in two clades on the COI tree.

thumbnail
Figure 3. Bayesian phylogenetic tree for combined COI and ITS2 sequence data for (A) Pontania and (B) Euura.

Outline of full midpoint-rooted tree is shown in the inset, host plants are indicated by individual label colours and symbols, and collection sites by symbol fills (see legend). Numbers above branches are posterior probabilities (%).

https://doi.org/10.1371/journal.pone.0116286.g003

Determinants of population-genetic structure

Within Pontania, pairwise COI ΦST values across population samples collected from different locations and hosts ranged from −0.187 to 0.987, and differentiation was statistically significantly non-zero (at P<0.05) for 83.33% of the compared pairs (S1 Table). ITS2-based Pontania ΦST values ranged from −0.098 to 1.000, and 84.84% of the contrasts were statistically significant (S2 Table). These results are mirrored in the Euura population samples (S1 and S2 Tables), in which pairwise COI ΦST = −0.277–1.000 and ITS2 ΦST = −0.116–0.959 (65.15% and 83.33% of pairs statistically significantly non-zero for COI and ITS2, respectively).

The hierarchical AMOVAs (Tables 2 and 3) show that, regardless of the order of the hierarchy, most of the observed genetic variation in both Pontania sequence datasets is explained by the host plant (COI: 50.47–68.66%; ITS2: 87.27–105.91%), while the effect of location is far weaker (COI: −11.10–11.05%; ITS2: −18.21–2.32%). The effect of host species is statistically significant in all cases, whereas the location effect is significantly non-zero only for the COI analysis in which locations are grouped within hosts (Table 2).

thumbnail
Table 2. Hierarchical analyses of molecular variance in Pontania and Euura gallers based on mitochondrial COI gene sequence data.

https://doi.org/10.1371/journal.pone.0116286.t002

thumbnail
Table 3. Hierarchical analyses of molecular variance in Pontania and Euura gallers based on nuclear ITS2 region sequence data.

https://doi.org/10.1371/journal.pone.0116286.t003

In analogous analyses based on the Euura samples, most of the observed genetic variation is again explained by host species (COI: 67.38–92.04%; ITS2: 61.63–72.96%; all P<0.05), while the influence of location is weak (COI: −9.75–16.72%; ITS2: −9.59–5.28%) and statistically significantly different from zero only when location is nested within host in the hierarchy (Tables 2 and 3).

Comparisons of differentiation in Pontania and Euura

In the COI dataset, average differentiation across hosts and locations (S1 Table) was statistically indistinguishable in the Euura (ΦST = 0.547±0.052 [mean ± s.e. mean]) and Pontania (ΦST = 0.538±0.035) samples (paired-samples t-test; t65 = −0.170, P = 0.865). However, for the ITS2 data (S2 Table), mean differentiation was significantly higher within Pontania (ΦST = 0.757±0.042) than in Euura (ΦST = 0.516±0.036) (paired-samples t-test; t65 = 5.359, P<0.001).

When populations from conspecific hosts in different locations were pooled, average COI differentiation across host species (Fig. 4A) did not differ between Euura (ΦST = 0.555±0.085) and Pontania (ΦST = 0.538±0.035) (paired-samples t-test; t14 = −0.186, P = 0.855). By contrast, mean differentiation in the ITS2 dataset (Fig. 4B) was significantly higher in Pontania (ΦST = 0.822±0.061) than in Euura (ΦST = 0.531±0.067) (paired-samples t-test; t14 = 3.321, P = 0.005).

thumbnail
Figure 4. Scatterplots of pairwise ΦST values estimated across pairs of willow host species in Pontania and Euura according to (A) COI and (B) ITS2 sequence data.

https://doi.org/10.1371/journal.pone.0116286.g004

Comparisons of ΦST values estimated across corresponding pairs of willow host species in the Euura and Pontania samples (Fig. 4) showed no evidence of a correlation in either the COI (Mantel test statistic r = −0.175, P = 0.256) or the ITS2 (r = 0.080, P = 0.393) datasets. Differentiation across hosts in the COI and ITS2 datasets correlated statistically significantly within Euura (Mantel test statistic r = 0.559, P = 0.022) but not within Pontania (Mantel test statistic r = 0.216, P = 0.209).

Discussion

One of the most central questions in evolutionary research concerns the repeatability of evolution [45]: if independent organismal lineages were to start from exactly the same conditions in identical ecological settings, would their evolutionary trajectories be similar? In research on replicated evolution, the main advantages of plant-feeding insects are their extreme species diversity and the fact that defining available niches (hosts) is easy [46]. For practical reasons, analyses of HAD and ecologically-based speciation in herbivores have thus far largely involved single insect taxa and their host plants (e.g., [7], [13], [47], [48]), but two recent studies have emphasized the need for broader analyses that search for commonalities across multiple herbivore groups: Stireman et al. [12] found evidence of HAD in four out of nine insect taxa that attack the North American goldenrod species Solidago altissima and S. gigantea, and Dickey and Medina [18] demonstrated that while the aphid Monelliopsis pecanis exhibits HAD across two broadly sympatric hickory species (Carya aquatica and C. illinoinensis), genetic structuring is essentially absent in the co-occurring tortricid moth Gretchena bolliana. Following the same logic, but on a larger scale, we estimated and compared the levels of HAD in two species-rich and closely related sawfly species complexes that exhibit wide overlap in their collective host repertoires. This approach makes it possible to infer how common HAD is and, more importantly, to study the reproducibility of genetic differentiation in similar environments.

Effects of geography and host-plant species on genetic differentiation in Pontania and Euura gallers

In addition to HAD, genetic divergence and speciation in insects could be driven by simple geographic isolation [17], [49]. An illuminating example in this respect is provided by the study by Craft et al. [50], who investigated DNA barcode variation in 28 lepidopteran (presumed) species collected from four plant families in eight locations scattered across the northern lowlands of New Guinea. They found substantial population subdivision in roughly half of the studied taxa: 14% showed evidence of HAD, 18% exhibited an isolation-by-distance pattern, 21% were differentiated by both hosts and geography, while the population structures of the remaining 46% were equivocal or essentially panmictic.

In our focal system, hierarchical AMOVAs (Tables 2 and 3) and pairwise ΦST estimates (S1 and S2 Tables) demonstrated frequent geographically-based population subdivision within both genera. However, the AMOVAs also showed that most of the variation in the mitochondrial and nuclear sequence datasets is explained by host species rather than by sampling locality; in the phylogenetic trees (Figs. 2 and 3), this is seen in that specimens tend to cluster according to willow species instead of collection site. Two locations were included in our analyses mainly in order to obtain independent replicate population samples within the studied taxa, and the relatively weak location effect was somewhat expected because the distance between Kilpisjärvi and Abisko is only slightly over 100 km; while this clearly exceeds the dispersal ability of individual sawflies, the homogeneous intervening landscape and vegetation present few obstacles to gene flow. However, both of the focal species-groups are distributed across the Holarctic region and individual species can have very broad ranges [24], [51], so stronger intraspecific spatial differentiation would be expected on larger geographical scales.

Genetic distances among clusters within the two genera are, however, very small, and most clades are to some degree intermixed with respect to host species, or at least include one or a few individuals from atypical hosts (Figs. 2 and 3). The absence of fixed diagnostic differences among the presumed species, and the lack of a correlation between COI and ITS2 differentiation within Pontania, could be explained by stochastic retention of ancestral polymorphisms in recently diverged lineages [52], [53]. In addition to coalescent stochasticity and incomplete lineage sorting, the observed shared polymorphisms and topological conflicts between the COI and ITS2 phylogenies could reflect occasional interbreeding among host races or incipient species that are still incompletely reproductively isolated from each other. Evidence for hybridization, and especially introgression of mtDNA haplotypes across species boundaries, has commonly been found in studies on genetic differentiation in insect herbivores [54][57]. Studies employing larger numbers of markers and individuals are needed to estimate the extent of hybridization in our focal species groups, but we note that occasional interbreeding among partly separated galler lineages seems probable in light of the recently-demonstrated widespread occurrence of interspecific gene flow in their Salix host plants [22].

Using allozyme-electrophoretic data, Nyman [27] showed that northern Fennoscandian bud-galling Euura lineages are divided into at least three mono- or oligophagous species. Our DNA sequence-based analyses support these results, especially with respect to the genetic distinctness of S. lapponum-associated samples (i.e., E. lappo Malaise), and the similarity of gallers on S. lanata and S. glauca (Fig. 3B). Individuals collected from the latter two willow species are essentially identical and intermixed in the trees, meaning that, like the earlier allozyme-based analyses, our results contradict Kopelke's [28] suggestion of the presence of two strict specialist species (E. lanatae and E. boreoalpina, respectively) on these presumedly distantly related [21] but chemically relatively similar [58] hosts. By contrast, our combined dataset supports Kopelke's [28] decision to elevate Malaise's [29]E. lappo var. hastatae’ to species rank (i.e., E. hastatae), although the species cannot be identified using COI data alone (Fig. 2B). The intriguing grouping of individuals collected from S. phylicifolia and S. myrsinifolia into two separate mixed clusters in the combined-data tree (Fig. 3B) seems to be driven mainly by the mitochondrial sequences (Fig. 2B), so we tentatively suggest that these samples represent a single lineage.

Host-based clustering is equally evident in the leaf-galling Pontania complex, in which especially samples collected from S. myrsinites and the arctic-alpine dwarf willow S. reticulata form distinct clades (Figs. 2 and 3; these clades correspond to the morphologically defined P. myrsiniticola Kopelke and P. reticulatae Malaise, respectively). The S. myrsinites-associated individual KV131 exhibited an atypical sequence in both COI and ITS2; therefore, it most likely indicates an oviposition error resulting in successful gall induction and larval development on a non-host willow. Pontania individuals sampled from the remaining Salix species are less consistently clustered according to their hosts, but the marked frequency differences among host-based population samples – resulting in statistically highly significant levels of genetic differentiation – indicate the presence of multiple independent lineages. However, our results cast doubt on the existence of separate leaf-galling species on S. myrsinifolia and S. borealis (P. varia and P. norvegica, respectively), as suggested by Kopelke [23]. In this respect, our results are in line with the interpretation that, rather than being a distinct species, S. borealis is a northern Fennoscandian high-altitude form or subspecies of the more widely distributed S. myrsinifolia (e.g., [43]).

The repeatability of HAD in gall-inducing sawflies

Phylogenetic studies have shown that, while the studied leaf-galling Pontania and bud-galling Euura groups are not sister taxa, they are of roughly equal age [25], [26]. Hence, it can be argued that the evolutionary opportunities and time available for genetic differentiation and speciation by specialization have been more or less equal in both taxa. Evolutionary repeatability should thus be manifest as comparable average levels of genetic subdivision across the studied hosts, but also as a correlation in the level of HAD across specific pairs of host species.

These predictions are not supported by our results: first, although we found clear evidence for frequent HAD in both galler taxa, average ΦST estimates across host species were similar only in the mitochondrial COI dataset, while differentiation in the nuclear ITS2 data was statistically significantly higher in Pontania than in Euura. One possible explanation for the discrepancy could be that the Pontania group is slightly older than the Euura group; if this is the case, our results would resemble the findings of Stireman et al. [12], who, in the aforementioned system including two Solidago species, found widely differing divergence times in Gnorimoschema moths, Rhopalomyia gall midges, and Eurosta tephritid flies. However, an alternative possibility is that the average rate of genetic divergence has been faster in Pontania. Second, lack of evolutionary replicability is indicated by the ‘shotgun’ scatterplots of across-host ΦST estimates in the two focal sawfly genera (Fig. 4). The disparate levels of differentiation suggest that although leaf- and bud-galling sawflies share a common taxonomic resource base, they differ in the way that they experience the adaptive landscape imposed by Salix species. This could, for example, be related to differences in the chemistry of leaves or buds, or to differences in the phenology of these tissues during the growing season. Bud gallers exhibit clear interspecific variation in average emergence time after pupation [59], and these differences are most likely linked to the phenology of their Salix hosts. In particular, differences in relative phenology across tissue types could lead to differential levels of gene flow among host-adapted insect lineages.

Conclusions

Studies on evolutionary radiations of taxa experiencing shared ‘adaptive landscapes’ can provide valuable insights into mechanisms that underlie ecological, phenotypic, and genetic divergence and/or convergence [60][63]. Our comparative analysis of genetic differentiation across shared Salix host species in northern Fennoscandian leaf-galling Pontania and bud-galling Euura sawflies revealed clear and frequent HAD in both species complexes. Nevertheless, mitochondrial and nuclear sequence variation is in most cases not diagnostic to samples collected from a given willow species, suggesting that, within the groups, species boundaries are still weakly defined, and that the identified lineages are positioned somewhere along the continuum between host races and fully independent species [10], [13], [64]. Like the few prior comparative analyses of HAD in multiple insect herbivore taxa [12], [18], [50] (see also [46]), our results suggest that speciation by ecological specialization plays a central role in herbivore diversification. However, the results have to be interpreted cautiously: like other endophagous insect groups, gall inducers are known to be unusually specialized in their host use [65], [66], meaning that the overall frequency of niche-driven speciation could be lower if other insect feeding guilds are also considered [12].

Despite genetic differentiation being present in both of the focal galler groups, we found differences in the overall level of HAD in the groups, and especially in the differentiation estimated across specific pairs of host species. This indicates that although the ecological selective background has been identical for both taxa, their evolutionary trajectories have been different. While the reasons for the disparities necessarily remain speculative, the few comparative studies done so far suggest that idiosyncrasies across herbivore taxa are likely to be a common feature in multispecies plant–herbivore systems [12], [50]. In this way, our results support Gould's [45] view that evolutionary trajectories are only weakly repeatable during replicated independent radiations.

Supporting Information

S1 Table.

Pairwise ΦST estimates among Pontania (below diagonal) and Euura (above diagonal) population samples collected from six different willow species in two locations, based on mitochondrial COI gene sequence data. Willow host names are abbreviated as follows: Smyr = S. myrsinifolia, Slap = S. lapponum, Sphy = S. phylicifolia, Sgla = S. glauca, Shas = S. hastata, Slan = S. lanata, collection sites are indicated by letters in parentheses (A = Abisko, K = Kilpisjärvi). All pairwise estimates are statistically significant at the P<0.05 level, except for values in parentheses = n.s.

https://doi.org/10.1371/journal.pone.0116286.s001

(DOCX)

S2 Table.

Pairwise ΦST estimates among Pontania (below diagonal) and Euura (above diagonal) population samples collected from six different willow species in two locations, based on nuclear ITS2 region sequence data. Willow host names are abbreviated as follows: Smyr = S. myrsinifolia, Slap = S. lapponum, Sphy = S. phylicifolia, Sgla = S. glauca, Shas = S. hastata, Slan = S. lanata, collection sites are indicated by letters in parentheses (A = Abisko, K = Kilpisjärvi). All pairwise estimates are statistically significant at the P<0.05 level, except for values in parentheses = n.s.

https://doi.org/10.1371/journal.pone.0116286.s002

(DOCX)

Acknowledgments

We thank the Kilpisjärvi and Abisko research stations for providing vital facilities during sample collection, Mia Valtonen for assistance in ARLEQUIN analyses, and Mikk Heidemaa and Marko Prous for useful tips on how to sequence ITS2 in sawflies. Constructive comments by six anonymous reviewers helped us to improve the manuscript.

Author Contributions

Conceived and designed the experiments: TN SL. Performed the experiments: SL KV. Analyzed the data: SL TN TM. Wrote the paper: SL TN TM.

References

  1. 1. Ehrlich PR, Raven PH (1964) Butterflies and plants: a study in coevolution. Evolution 18:586–608.
  2. 2. Strong DR, Lawton JH, Southwood TRE (1984) Insects on plants: community patterns and mechanisms. Cambridge, MA: Harvard University Press.
  3. 3. Janz N (2011) Ehrlich and Raven revisited: mechanisms underlying codiversification of plants and enemies. Annu Rev Ecol Evol Syst 42:71–89.
  4. 4. Berlocher SH, Feder JL (2002) Sympatric speciation in phytophagous insects: moving beyond controversy? Annu Rev Entomol 47:773–815.
  5. 5. Agosta SJ (2006) On ecological fitting, plant-insect associations, herbivore host shifts, and host plant selection. Oikos 114:556–565.
  6. 6. Matsubayashi KW, Ohshima I, Nosil P (2010) Ecological speciation in phytophagous insects. Entomol Exp Appl 134:1–27.
  7. 7. Via S, Bouck AC, Skillman S (2000) Reproductive isolation between divergent races of pea aphids on two hosts. II. Selection against migrants and hybrids in the parental environments. Evolution 54:1626–1637.
  8. 8. Funk DJ, Filchak KE, Feder JL (2002) Herbivorous insects: model systems for the comparative study of speciation ecology. Genetica 116:251–267.
  9. 9. Nyman T (2010) To speciate, or not to speciate? Resource heterogeneity, the subjectivity of similarity, and the macroevolutionary consequences of niche-width shifts in plant-feeding insects. Biol Rev 85:393–411.
  10. 10. Drès M, Mallet J (2002) Host races in plant-feeding insects and their importance in sympatric speciation. Philos Trans R Soc Lond B Biol Sci 357:471–492.
  11. 11. Egan SP, Funk DJ (2009) Ecologically dependent postmating isolation between sympatric host forms of Neochlamisus bebbianae leaf beetles. Proc Natl Acad Sci USA 106:19426–19431.
  12. 12. Stireman JO, Nason JD, Heard SB (2005) Host-associated genetic differentiation in phytophagous insects: general phenomenon or isolated exceptions? Evidence from a goldenrod-insect community. Evolution 59:2573–2587.
  13. 13. Peccoud J, Ollivier A, Plantegenest M, Simon J-C (2009) A continuum of genetic divergence from sympatric host races to species in the pea aphid complex. Proc Natl Acad Sci USA 106:7495–7500.
  14. 14. Rundle HD, Nosil P (2005) Ecological speciation. Ecol Lett 8:336–352.
  15. 15. Winkler IS, Mitter C (2008) The phylogenetic dimension of insect-plant interactions: a review of recent evidence. In: Specialization, speciation, and radiation: the evolutionary biology of herbivorous insects. Berkeley, CA: University of California Press. pp. 240–263.
  16. 16. Nyman T, Vikberg V, Smith DR, Boevé JL (2010) How common is ecological speciation in plant-feeding insects? A ‘Higher’ Nematinae perspective. BMC Evol Biol 10:266.
  17. 17. Imada Y, Kawakita A, Kato M (2011) Allopatric distribution and diversification without niche shift in a bryophyte-feeding basal moth lineage (Lepidoptera: Micropterigidae). Proc R Soc B Biol Sci 278:3026–3033.
  18. 18. Dickey AM, Medina RF (2010) Testing host-associated differentiation in a quasi-endophage and a parthenogen on native trees. J Evol Biol 23:945–956.
  19. 19. Dickey AM, Medina RF (2012) Host-associated genetic differentiation in pecan leaf phylloxera. Entomol Exp Appl 143:127–137.
  20. 20. Argus GW (1997) Infrageneric classification of Salix (Salicaceae) in the New World. Syst Bot Monogr 52:1–121.
  21. 21. Skvortsov AK (1999) Willows of Russia and adjacent countries. Joensuu: University of Joensuu.
  22. 22. Percy DM, Argus GW, Cronk QC, Fazekas AJ, Kesanakurti PR, et al. (2014) Understanding the spectacular failure of DNA barcoding in willows (Salix): does this result from a trans-specific selective sweep? Mol Ecol 23:4737–4756.
  23. 23. Kopelke JP (1999) Gallenerzeugende Blattwespen Europas: Taxonomische Grundlagen, Biologie und Ökologie (Tenthredinidae: Nematinae: Euura, Phyllocolpa, Pontania). Cour Forsch–Inst Senckenb 212:1–183.
  24. 24. Roininen H, Nyman T, Zinovjev A (2005) Biology, ecology, and evolution of gall-inducing sawflies (Hymenoptera: Tenthredinidae and Xyelidae). In: Raman A, Schaefer CW, Withers TMeditors. Biology, ecology and evolution of gall-inducing arthropods, Volume 1 and 2. Enfield, NH: Science Publishers, Inc. pp. 467–494.
  25. 25. Nyman T, Widmer A, Roininen H (2000) Evolution of gall morphology and host-plant relationships in willow-feeding sawflies (Hymenoptera: Tenthredinidae). Evolution 54:526–533.
  26. 26. Nyman T, Bokma F, Kopelke JP (2007) Reciprocal diversification in a complex plant–herbivore–parasitoid food web. BMC Biol 5:49.
  27. 27. Nyman T (2002) The willow bud galler Euura mucronata Hartig (Hymenoptera: Tenthredinidae): one polyphage or many monophages? Heredity 88:288–295.
  28. 28. Kopelke JP (2001) The species groups of Euura mucronata and E. laeta in Europe (Insecta, Hymenoptera, Tenthredinidae, Nematinae). Senckenb Biol 81:191–225.
  29. 29. Malaise R (1920) Beiträge zur Kenntnis schwedischer Blattwespen. Entomol Tidskr 40:97–128.
  30. 30. Nyman T, Zinovjev A, Vikberg V, Farrell B (2006) Molecular phylogeny of the sawfly subfamily Nematinae (Hymenoptera: Tenthredinidae). Syst Entomol 31:569–583.
  31. 31. Heidemaa M, Nuorteva M, Hantula J, Saarma U (2004) Dolerus asper Zaddach, 1859 and Dolerus brevicornis Zaddach, 1859 (Hymenoptera: Tenthredinidae), with notes on their phylogeny. Eur J Entomol 101:637–650.
  32. 32. Magnacca KN, Brown MJ (2010) Mitochondrial heteroplasmy and DNA barcoding in Hawaiian Hylaeus (Nesoprosopis) bees (Hymenoptera: Colletidae). BMC Evol Biol 10:174.
  33. 33. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, et al. (2007) Clustal W and Clustal X version 2.0. Bioinformatics 23:2947–2948.
  34. 34. Maddison WP, Maddison DR (2010) Mesquite: a modular system for evolutionary analysis. Version 2.74. Available at: http://mesquiteproject.org/mesquite/download/download.html.
  35. 35. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, et al. (2012) MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol 61:539–542.
  36. 36. Miller MA, Pfeiffer W, Schwartz T (2010) Creating the CIPRES Science Gateway for inference of large phylogenetic trees. In: Proceedings of the Gateway Computing Environments Workshop (GCE). 14 Nov. 2010, New Orleans, LA. pp. 1–8.
  37. 37. Simmons MP, Ochoterena H (2000) Gaps as characters in sequence-based phylogenetic analyses. Syst Biol 49:369–381.
  38. 38. Borchsenius F (2009) FastGap 1.2. Available at: http://www.aubot.dk/FastGap_home.htm.
  39. 39. Rambaut A, Suchard M, Drummond AJ (2013) Tracer v1.6.0. Available at: http://beast.bio.ed.ac.uk/Tracer.
  40. 40. Excoffier L, Lischer HE (2010) Arlequin suite ver 3.5: a new series of programs to perform population genetics analyses under Linux and Windows. Mol Ecol Resour 10:564–567.
  41. 41. Excoffier L, Smouse PE, Quattro JM (1992) Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics 131:479–491.
  42. 42. Darriba D, Taboada GL, Doallo R, Posada D (2012) jModelTest 2: more models, new heuristics and parallel computing. Nat Methods 9:772–772.
  43. 43. Hämet-Ahti L, Suominen J, Ulvinen T, Uotila P (1998) Retkeilykasvio (Field flora of Finland). Helsinki: Finnish Museum of Natural History, Botanical Museum.
  44. 44. McCune B, Mefford MJ (2006) PC-ORD. Multivariate analysis of ecological data. Version 5. MjM Software Design, Gleneden Beach, Oregon, USA.
  45. 45. Gould SJ (1989) Wonderful life: the Burgess shale and the nature of history. New York: W. W. Norton & Co.
  46. 46. Egan SP, Hood GR, DeVela G, Ott JR (2013) Parallel patterns of morphological and behavioral variation among host-associated populations of two gall wasp species. PLoS ONE 8:e54690.
  47. 47. Dorchin N, Scott ER, Clarkin CE, Luongo MP, Jordan S, Abrahamson WG (2009) Behavioural, ecological and genetic evidence confirm the occurrence of host-associated differentiation in goldenrod gall-midges. J Evol Biol 22:729–739.
  48. 48. Hernández-Vera G, Mitrović M, Jović J, Tosevski I, Caldara R, et al. (2010) Host-associated genetic differentiation in a seed parasitic weevil Rhinusa antirrhini (Coleptera: Curculionidae) revealed by mitochondrial and nuclear sequence data. Mol Ecol 19:2286–2300.
  49. 49. Aoki K, Kato M, Murakami N (2011) Phylogeography of phytophagous weevils and plant species in broadleaved evergreen forests: a congruent genetic gap between western and eastern parts of Japan. Insects 2:128–150.
  50. 50. Craft KJ, Pauls SU, Darrow K, Miller SE, Hebert PD, et al. (2010) Population genetics of ecological communities with DNA barcodes: an example from New Guinea Lepidoptera. Proc Natl Acad Sci USA 107:5041–5046.
  51. 51. Smith DR (1979) Suborder Symphyta. In: Krombein KV, Hurd PD, Smith DR, Burks BD, editors. Catalog of Hymenoptera in America North of Mexico. Washington, DC: Smithsonian Institution Press. pp. 3–137.
  52. 52. Pamilo P, Nei M (1988) Relationships between gene trees and species trees. Mol Biol Evol 5:568–583.
  53. 53. Nowell RW, Charlesworth B, Haddrill PR (2011) Ancestral polymorphisms in Drosophila pseudoobscura and Drosophila miranda. Genet Res 93:255–263.
  54. 54. Linnen CR, Farrell BD (2007) Mitonuclear discordance is caused by rampant mitochondrial introgression in Neodiprion (Hymenoptera: Diprionidae) sawflies. Evolution 61:1417–1438.
  55. 55. Gompert Z, Forister ML, Fordyce JA, Nice CC (2008) Widespread mito-nuclear discordance with evidence for introgressive hybridization and selective sweeps in Lycaeides. Mol Ecol 17:5231–5244.
  56. 56. Marshall DC, Hill KB, Cooley JR, Simon C (2011) Hybridization, mitochondrial DNA phylogeography, and prediction of the early stages of reproductive isolation: lessons from New Zealand cicadas (genus Kikihia). Syst Biol 60:482–502.
  57. 57. Mardulyn P, Othmezouri N, Mikhailov YE, Pasteels JM (2011) Conflicting mitochondrial and nuclear phylogeographic signals and evolution of host-plant shifts in the boreo-montane leaf beetle Chrysomela lapponica. Mol Phylogenet Evol 61:686–696.
  58. 58. Julkunen-Tiitto R (1989) Phenolic constituents of Salix: a chemotaxonomic survey of further Finnish species. Phytochemistry 28:2115–2125.
  59. 59. Nyman T (2000) Phylogeny and ecological evolution of gall-inducing sawflies (Hymenoptera: Tenthredinidae). PhD Thesis, University of Joensuu, Finland.
  60. 60. Rüber L, Verheyen E, Meyer A (1999) Replicated evolution of trophic specializations in an endemic cichlid fish lineage from Lake Tanganyika. Proc Natl Acad Sci USA 96:10230–10235.
  61. 61. Losos JB (2010) Adaptive radiation, ecological opportunity, and evolutionary determinism. Am Nat 175:623–639.
  62. 62. Burbrink FT, Chen X, Myers EA, Brandley MC, Pyron RA (2012) Evidence for determinism in species diversification and contingency in phenotypic evolution during adaptive radiation. Proc R Soc B Biol Sci 279:4817–4826.
  63. 63. Mahler DL, Ingram T, Revell LJ, Losos JB (2013) Exceptional convergence on the macroevolutionary landscape in island lizard radiations. Science 341:292–295.
  64. 64. Powell TH, Hood GR, Murphy MO, Heilveil JS, Berlocher SH, et al. (2013) Genetic divergence along the speciation continuum: the transition from host race to species in Rhagoletis (Diptera: Tephritidae). Evolution 67:2561–2576.
  65. 65. Hardy NB, Cook LG (2010) Gall-induction in insects: evolutionary dead-end or speciation driver? BMC Evol Biol 10:257.
  66. 66. Novotny V, Miller SE, Baje L, Balagawi S, Basset Y, et al. (2010) Guild-specific patterns of species richness and host specialization in plant–herbivore food webs from a tropical forest. J Anim Ecol 79:1193–1203.