TMCnet News

Gene-Flow in a Mosaic Hybrid Zone: Is Local Introgression Adaptive? [Genetics]
[July 31, 2014]

Gene-Flow in a Mosaic Hybrid Zone: Is Local Introgression Adaptive? [Genetics]


(Genetics Via Acquire Media NewsEdge) ABSTRACT Genome-wide scans of genetic differentiation between hybridizing taxa can identify genome regions with unusual rates of introgression. Regions of high differentiation might represent barriers to gene flow, while regions of low differentiation might indicate adaptive introgression-the spread of selectively beneficial alleles between reproductively isolated genetic backgrounds. Here we conduct a scan for unusual patterns of differentiation in a mosaic hybrid zone between two mussel species, Mytilus edulis and M. galloprovincialis. One outlying locus, mac-1, showed a characteristic footprint of local introgression, with abnormally high frequency of edulis-derived alleles in a patch of M. galloprovincialis enclosed within the mosaic zone, but low frequencies outside of the zone. Further analysis of DNA sequences showed that almost all of the edulis allelic diversity had introgressed into the M. galloprovincialis background in this patch. We then used a variety of approaches to test the hypothesis that there had been adaptive introgression at mac-1. Simulations and model fitting with maximum-likelihood and approximate Bayesian computation approaches suggested that adaptive introgression could generate a "soft sweep," which was qualitatively consistent with our data. Although the migration rate required was high, it was compatible with the functioning of an effective barrier to gene flow as revealed by demographic inferences. As such, adaptive introgression could explain both the reduced intraspecific differentiation around mac-1 and the high diversity of introgressed alleles, although a localized change in barrier strength may also be invoked. Together, our results emphasize the need to account for the complex history of secondary contacts in interpreting outlier loci.



(ProQuest: ... denotes formulae omitted.) GENETIC barriers between related lineages are often semipermeable, varying in strength across the genome (Harrison 1986). As such, hybridization can lead to adaptive introgression, the transfer of selectively beneficial alleles between species (Parsons et al. 1993; Arnold 2004; Helico- nius Genome Consortium 2012; Hedrick 2013). However, the frequency and importance of such hybridization events remain hotly debated (Mallet 2007; Abbott et al. 2013; Barton 2013). It is now common to address such questions by scan- ning genome sequences for regions of enhanced or reduced differentiation. Universally advantageous alleles can usually cross barriers to gene flow without much delay (e.g., Piálek and Barton 1997), and neutral loci linked to such alleles are expected to display unusually low levels of differentiation. In contrast, barrier loci will delay the introgression of neutral alleles in proportion to their linkage (Barton 1979; Barton and Bengtsson 1986); that is the basis for the investigation of genomic islands of differentiation (Turner et al. 2005; Hohenlohe et al. 2010; Nadeau et al. 2012).

Interpreting regions of unusual differentiation can be difficult, not only because of variation in recombination across the genome (Nachman and Payseur 2012; Roesti et al. 2013), but also because reduced differentiation might be caused by processes other than adaptive introgression. In particular, pop- ulation history may promote multiple contacts between the same lineages in different locations, and so the resulting bar- riers to gene flow might vary from place to place, according to local ecological gradients and population connectivity. In ac- cordance with this hypothesis, reversed or modified associa- tions between genetic differentiation and habitat variation (Bierne et al. 2011; Jackson et al. 2012) and partial parallel- ism in genomic divergence (Gagnaire et al. 2013) have both been observed in replicated contact zones of the same species pair. Differentiating between these hypotheses is challenging, because adaptive introgression leaves a genomic signature that can be substantially different from the classic "hard sweep" (Maynard Smith and Haigh 1974) where adaptive substitution is associated with a single new mutation. Adaptive introgres- sion may involve multiple migrant copies of the same beneficial allele (Pennings and Hermisson 2006) leading to a "soft-sweep" signature, which can be difficult to discriminate from a local increase in the neutral introgression rate (see Table 1).


Here we present a multilocus scan for local introgression between the marine mussels Mytilus edulis and M. gallopro- vincialis. These species appear to have had a complex history of fragmentation and colonization during the Quaternary, as is the case for almost all temperate organisms (Hewitt 2011). Recent analyses suggest that M. edulis and M. galloprovincialis diverged ^2.5 million years (MY), followed by a secondary contact beginning about 0.7 MY (Roux et al. 2014). Today, they are isolated by multifarious pre- and postzygotic mechanisms (Bierne et al. 2002, 2003a, 2006). Due to natural replications of contact zones, the Mytilus species complex is particularly suit- able for studying the different possible outcomes of secondary contact. Here we study an original mosaic hybrid zone in Europe, extending from the Mediterranean Sea to the North Sea (Bierne et al. 2003b; Hilbish et al. 2012) and to the British Isles (Skibinski et al. 1983). We focus on the French section of the mosaic zone that includes two patches per species, one peripheral and one enclosed within the zone (Figure 1).

To study introgression between these populations, we used previously published data at 440 loci (422 amplified fragment length polymorphism, AFLP, and 18 codominant nuclear markers; Gosset and Bierne 2012). Initial analyses identified a nuclear marker, mac-1, with high interspecific differentiation in the peripheral patch but an anomalously low level of differentiation in the enclosed patch in Brittany. This outlier locus was therefore a candidate for adaptive introgression in a geographically localized region.

To test the plausibility of this hypothesis, we first sequenced a 3.1-kb region around mac-1 and analyzed how allele fre- quency varies along the chromosomal region. Unlike another M. edulis locus analyzed previously (Bierne 2010), we failed to identify variation in differentiation on such a small chromo- somal scale. We then determined how well our data might be explained by a model of adaptive introgression (Pennings and Hermisson 2006). Inferences indicated that the level of gene flow implied by the adaptive introgression hypothesis was high, but still compatible with the strength of the genetic bar- rier between the two species. We further propose that the barrier to gene flow in the vicinity of mac-1 could be simply more permeable in certain hybrid zones of the mosaic. Overall, we show that although the molecular signature of adaptive introgression is difficult to identify, dedicated methods that account for the history of speciation can help in interpreting outlying levels of introgression.

Materials and Methods Study sites and sampling The zone of study, along the European coast, is characterized by three successive transitions between panmictic patches of each species (Figure 1; Bierne et al. 2003c; Hilbish et al. 2012). We used two geographical samples of M. edulis:the peripheral patch of the North Sea (Wadden Sea, Holland); and the enclosed patch of the Bay of Biscay (Lupin, France). We also used two geographical samples of M. galloprovincialis: the peripheral patch of the Iberian Coast (Faro, Portugal); and the enclosed patch of Brittany (Roscoff, France). These four samples have been described in Faure et al. (2008) and have been established to be representative of monospecific pan- mictic patches. In addition, we used a sample of M. trossulus (Tadoussac, Canada) to serve as an outgroup. Forty-eight individuals per sample were examined, except for the Brit- tany sample, which comprised 87 individuals. Genomic DNA was extracted from adults using the DNeasy Blood and Tissue Kit (Qiagen) following the manufacturers protocol.

Multilocus scan For the multilocus scan, we combined existing data, comprising 422 AFLP markers, 13 codominant nuclear markers (Gosset and Bierne 2012), and 5 allozyme markers (Faure et al. 2008). To combine loci with various level of diversity, nuclear codom- inant markers were transformed into biallelic loci by pooling alleles according to their frequencies in the M. galloprovincialis and M. edulis reference samples (see McDonald 1994). AFLP markers are subject to several well-known caveats, notably fragment-size homoplasy (Caballero et al. 2008; Whitlock et al. 2008). To reduce this problem, we excluded AFLP bands smaller than 50 bp, which are most prone to homoplasy (Vekemans et al. 2002). Allele frequencies of AFLP markers were esti- mated with the Bayesian method of Zhivotovsky (1999) in AFLPsurv v. 1.0 (Vekemans et al. 2002). All data were com- bined for the outlier tests.

To ensure robustness, and following the recommendation of Pérez-Figueroa et al. (2010), we used six distinct methods to identify loci with unusual levels of genetic differentiation. These are the methods of Lewontin and Krakauer (1973), Beaumont and Nichols (1996), Vitalis et al. (2003), Beaumont and Balding (2004), Bonhomme et al. (2010), and a custom simulation method in which the neutral envelope of FST is obtained from simulations with the parameter estimates of the best-supported demographic model. Full details are given in Supporting Information, File S1.

Data analyses at the outlier locus mac-1 Chromosomal walking along mac-1: We sequenced a re- gion of 3.1 kb around mac-1 taking advantage of published sequences (Daguin et al. 2001), an additional upstream se- quence (M. Ohresser, personal communication) and looking for homology between coding regions and expressed sequence tags of Mytilus (Tanguy et al. 2008). To describe the local variation of the edulis allele frequency around mac-1,we walked along the sequence and looked for evenly spaced genetic polymorphisms (see Figure S2 for a diagram). Four polymorphisms, two PCR product-length polymorphisms and two SNPs, were chosen, based on the allelic genealogies, to be discriminant between the two species. We extensively genotyped our four samples (North Sea, Brittany, Bay of Biscay, and Iberian Coast) for these four polymorphisms. See File S1 for technical details.

Genetic analysis: For the PCR product-length polymorphisms described above, we computed classical genetic statistics on the subset of edulis alleles (i.e., alleles that group within the diversity of M. edulis within the mac-1 genealogy; see Figure 3) using Genetix 4.05.2 (Belkhir et al. 1996-2004): the num- berofallelicclasses(k), the heterozygosity (He, Nei 1978), departure from Hardy-Weinberg equilibrium, and genetic dif- ferentiation were estimated by f and Q (Weir and Cockerham 1984), respectively, and significance was tested with 1000 permutations. Finally, we evaluated the departure from the neutral expectation at mutation-drift equilibrium with the Ewens test (e, Slatkin 1994; Slatkin 1996).

Sequence alignment was performed with Multalin (Corpet 1988) and verified by eye in BioEdit 7.5.3 (Hall 1999). Align- ment gaps were excluded from further analyses. We inferred allelic genealogies for the three regions sequenced using the neighbor-joining algorithm implemented in Mega 5.0 (Kumar et al. 2004). We rooted the allele genealogy with the outgroup M. trossulus. We computed classical genetic statistics on DNA sequences using DNAsp (Rozas et al. 2003): the number of polymorphic sites (S), the number of synonymous (Ss), non- synonymous (Sns) and noncoding (Snc) mutations, levels of nucleotide diversity estimated from the number of polymor- phic sites (uS, Watterson 1975) and from pairwise differences (up, Tajima 1983) and the minimum number of recombination events (Rm, Hudson and Kaplan 1985). We also computed two indicators of the distortion of the allele-frequency spec- trum from the neutral mutation-drift equilibrium expecta- tion: (1) Tajima's D (Tajima 1989a; 1989b) and (2) Fay and Wu's H (Fay and Wu 2000); significant departure from stan- dard neutral model were assessed using standard-coalescent simulations.

Model of adaptive introgression: To assess the hypothesis of adaptive introgression at mac-1, we considered a model of adaptive introgression via recurrent migration, introduced by Pennings and Hermisson (2006).

This model considers a universally beneficial allele that introgresses from the donor population(s) to the recipient population via recurrent migration and then sweeps to global fixation. Formally, the model considers two Wright-Fisher populations, each of N individuals. Each individual is char- acterized by two completely linked haploid loci. The first locus is biallelic, with alleles B and b. The B allele has a se- lective advantage s over the b allele. The second, linked locus, is neutral with an infinite number of possible alleles, and new mutations arising at rate m (the per-gene mutation rate). The donor population is initially fixed for the benefi- cial B allele, while the recipient population is initially fixed for the alternative b allele, and both populations are as- sumed to have reached mutation-drift equilibrium at the neutral locus. After equilibration follows a period of second- ary contact in which the populations exchange migrants at rate m. During this secondary contact, the beneficial B allele is introduced into the recipient population, where it will even- tually reach fixation. The three most important parameters of this model are the scaled quantities u =2Nm, M =2Nm,and a =2Ns (Pennings and Hermisson 2006).

To fit this model to data, we used the mac-1 alleles sam- pled from the donor populations (the two M. edulis patches) and the recipient population (the Brittany patch of M. gallo- provincialis). Analytical results assume complete linkage and recent completion of the sweep (see below), and so, to make our inferences robust to the violation of these assumptions, we take as data only those M. galloprovincialis alleles that group within the diversity of M. edulis within the genealogy (i.e., only the alleles denoted with open circles in the edulis clade, Figure 3).

Maximum-likelihood approach: We first fit the data using a maximum-likelihood (ML) approach. Our data in this case comprised the number of alleles sampled in the donor and recipient populations (nd and nr) and the number of distinct allelic classes in these samples (kd and kr). Under the assump- tions described above, the probability of observing the data from the donor population is given by the Ewens' sampling formula (Ewens 1972), ... (1) where ... is an unsigned Stirling number of the first kind, i.e., the number of permutations of n elements in k disjoint cycles (Charalambides and Singh 1988). Pennings and Hermisson (2006) showed that adaptive introgression has some formal similarities to Equation 1. In particular, if a = 2Ns ^ 1, then after the beneficial allele has reached fixation in the donor population, the number of distinct immigrant alleles that contribute to the sweep also approximate Ewens' distribution, but with the scaled migration rate, M,playingthe roleofthescaledmutationrate,u (Pennings and Hermisson 2006). We cannot observe the number of immigrant alleles directly, but we do know that these migrants were drawn at random from the donor population and that the diversity in the donor population is described by Ewens' distribution with parameter u. As such, the likelihood of observing our data are well approximated by ... (2) (where the summation is over the number of possible migrants from the donor population, which appear in the sample from the recipient population). Note that the approximation of Equation 2 does not depend explicitly on the strength of selection, a, but this new result does allow us to estimate the parameters u and M from our mac-1 data. The parame- ters that maximize Equation 2, u^ and M^ , are the maximum- likelihood estimates. Confidence intervals on a given parameter were the values that reduced the log likelihood by two units- obtained by maximizing the likelihood, conditional on the parameter of interest taking a suboptimal value (Edwards 1992). An R script that implements Equation 2 is included as File S2.

Approximate Bayesian computation approach: Equation 2 is an approximation (Pennings and Hermisson 2006); fur- thermore, it considers only the number of allelic classes (kd and kr) and, therefore, ignores some of the information in the data, such as the allele-frequency spectrum. Accordingly, we also estimated model parameters using an approximate Bayesian computation (ABC) approach (Beaumont et al. 2002). This approach uses forward simulation of the adaptive intro- gression model and then compares summary statistics of the simulated and observed data.

To simulate the model, mutation-drift equilibrium was first achieved by sampling directly from the full Ewens (1972) distribution, as implemented in the program MONTECARLO (Slatkin 1994). Then, we ran forward-in-time simulations of the secondary contact, ending each simulation when the ben- eficial allele had reached fixation in the recipient population.

To compare the fit of the simulations to our data, we used several summary statistics. First, for each of the two species where d denotes the donor species and r denotes the recipient species, we calculated (1) the number of allelic classes (kd and kr), (2) the expected heterozygosity at the locus (Hed and Her; Nei 1978), (3) the standard deviation in frequencies over all allelic classes (Sdd and Sdr), and (4) the proportion of private alleles (not found in the other species, %Pd and % Pr). We also used (5) the between-species diversity FST (Nei 1973) and (6) the proportional diversity after introgression described by Her/Hed. All of these summary statistics for the true data are given in Table S1.

To estimate the parameters, we performed 950,000 for- ward simulations in total. For each simulation, parameters were log-tangent transformed (Hamilton et al. 2005), and we then calculated the Euclidean distance between the observed and simulated transformed parameters. The 1000 simulations with the smallest associated Euclidean distance were then retained, and the posterior distribution of the scaled param- eters was estimated by means of weighted nonlinear multi- variate regressions of the parameters on the summary statistics (Blum and Francois 2010). For each regression, 50 feed-forward neural networks and 25 hidden networks were trained using the R package abc (Csilléry et al. 2012) and results were averaged over the replicate networks. Priors on the model parameters were all uniform (flat), and used the following ranges: u ^ [0.2, 40], M ^ [0.2, 200], a ^ [2, 2000].

We evaluated the performance of the ABC method in two ways. First, we performed a goodness-of-fit test (Cornuet et al. 2010). This involved simulating 50,000 data sets with parameters drawn from their estimated posterior distributions and then comparing the summary statistics of these simulated data sets to the statistics of the observed data. Second, we computed the mean bias statistic, b ¼ 1=n i ½ei =vi ^,wheren is the number of summary statistics, ei is the median estimation, and vi is the true median value of the ith data set (Excoffier et al. 2005). We generated 100 pseudo-observed data sets with known parameter values drawn from the prior distri- butions and computed b for each scaled parameter.

Forward simulations of adaptive introgression written in C++ are available (see File S2) together with R scripts implementing a simple version of the ABC approach.

Models of speciation: To compare our estimate of M at mac- 1 to the extent of gene fl ow between the Brittany patch of M. galloprovincialis and M. edulis, we took advantage of the ABC approach described in Roux et al. (2013) that accounts for a putative genome-wide heterogeneity in introgression rates. We obtained DNA sequence data at the eight nuclear loci described in Roux et al. (2014) for the North Sea patch of M. edulis and the Brittany patch of M. galloprovincialis. Only silent positions (i.e., synonymous polymorphisms in coding regions and noncoding polymorphisms in introns or intergenic regions) were retained in the analysis.

We investigated four models of speciation differentiated by their temporal patterns of introgression (see Figure S8): (1) strict isolation (SI) between the two daughter species, (2) isolation with migration (IM) assuming continuous gene flow since the two species started to diverge, (3) ancient migration (AM), where migration is restricted to the early period of speciation, and (4) secondary contact (SC), where the two daughter populations first evolve in strict isolation and then experience gene flow in a secondary contact. For models with gene flow (IM, AM, and SC), we compared two alternative scenarios in which the effective migration rate was either homogeneous or heterogeneous among loci.

Five million multilocus simulations were performed for each model using the coalescent simulator Msnsam (Hudson 2002; Ross-Ibarra et al. 2008). We then compared the simu- lated and observed data sets by using an array of summary statistics widely used in the literature (Wakeley and Hey 1997; Fagundes et al. 2007; Ross-Ibarra et al. 2008): (1) nucleotide diversity (p, Tajima 1983); (2) Watterson's uw (Watterson 1975); (3) net interspecific divergence (netdivedu-gal); and (4) between-species differentiation computed as 1 2 pS/pT, where pS is the average pairwise nucleotide diversity within species and pT is the total pairwise nucleotide diversity of the pooled sample across species. We assessed departure from mutation-drift equilibrium using Tajima's D (Tajima 1989a, b). We also classified the variable genomic positions accord- ing to whether they are exclusively polymorphic in M. edulis (Sx.edu), exclusively polymorphic in M. galloprovincialis (Sx.gal), with different alleles fixed in both species (Sf), or sharing the same polymorphism (SS). We computed the average and stan- dard deviation of each of these statistics across the surveyed loci by using MScalc (available from http://www.abcgwh. sitew.ch/;seeRouxet al. 2011). We provide the observed values computed from the sequenced data in Table S2.Full detail of priors, model choice and model checking procedures, as well as parameter estimation, are described in File S1.

Results Detection of interspecific and intraspecific outliers Figure 2A compares interspecific and intraspecific differen- tiation at screened loci. The corresponding figure in M. edulis is provided (see Figure S1). Differentiation between patches of the same species was generally low (in M.gallo- provincialis FST = 0.021; in M. edulis FST = 0.024), but highly variable between loci, and a few loci were identified as outliers using several methods of detection. In particular, all methods agreed that three loci were outliers within M. galloprovincialis (ACG.CGA.206 FST = 0.189; mac-1 FST = 0.192 and CAG.CTC.317 FST = 0.224; Figure 2A), and two loci within M. edulis (ACG.CGA.152 FST = 0. 294 and CAG. CGA.363 FST = 0.374; Figure S1). In the comparison be- tween species, differentiation was also low (FST = 0.068) but the distribution of FST values was overdispersed, as one would expect when a semipermeable barrier to gene flow isolates two species. The locus mac-1, which was highly differentiated within M. galloprovincialis, was also highly differentiated between species (FST = 0.507) and was the only locus to be detected as an outlier in both kinds of comparison.

To see this in more detail, Figure 2B shows allele frequen- cies in all four patches at the three loci identified as outliers within M. galloprovincialis. While all three loci show allele- frequency differences within M. galloprovincialis, only mac-1 shows allele-frequency differences between the Brittany patch of M. galloprovincialis and the neighboring patches of M. edulis.

Detailed analysis of mac-1 Because the pattern of differentiation at mac-1 was unique, this locus was investigated in more detail. Specifically, we sequenced three regions, comprising a total of 3.1 kb in the vicinity of mac-1 (Figure S2). All three regions contained several SNPs, and the first region also contained indels that were used to define PCR primers that generate length poly- morphism of the PCR product, indel-in0. Summary statistics of the genetic diversity within each sequenced region are found in Table 2, while Table 3 describes the genetic diver- sity of the two PCR product-length polymorphisms.

We next constructed genealogies of the three regions. Figure 3 shows the genealogy of region 2 where alleles from the Brittany patch of M. galloprovincialis are found in two distinct clades, grouping with (1) alleles from both M. edulis patches ("edulis clade"; Figure 3) and (2) with the other galloprovincialis alleles ("galloprovincialis clade"; Figure 3). Similar patterns are also found in regions 1 and 3 (Figure S3) and were robust to the exclusion of possible recombi- nants (Table 2).

Together, these patterns are consistent with the results of FST outlier tests and suggest that there has been an intro- gression of mac-1 alleles from M. edulis into the Brittany patch of M. galloprovincialis, which lies between them. Fur- thermore, comparison of the Brittany alleles that group in the edulis clade (open circles within the "edulis clade"; Fig- ure 3), with the related alleles found in the two M. edulis patches (solid squares and circles, Figure 3), suggests that very similar levels of genetic diversity are found in all three patches (see Table 2 and Table 3). For example, the pro- portion of diversity introgressed at the locus indel-in0 is Hegal/Heedu = 0.67 and at mac-1 is Hegal/Heedu = 0.96 (see Table 3). Additionally, there is no evidence of departure from neutrality in any of the three sets of alleles (see the results of Tajima's D and Fay and Wu's H in Table 2).

One possible interpretation of these results is that mac-1 has been subject to adaptive introgression from M. edulis into Brittany patch of M. galloprovincialis. The following sections use several methods to test the validity of this hypothesis.

Testing for the chromosomal signature of adaptive introgression One genomic signature of local adaptation is a single sharp peak in the frequencies of linked neutral alleles, centered on the adaptive polymorphism (Charlesworth et al. 1997). To test for this signature we identified four polymorphisms, two indels and two SNPs, that mapped to an internal branch of their corresponding genealogy (Figure 3 for region 2 and Figure S3 for regions 1 and 3). These polymorphisms were evenly spaced along the 3.1-kb sequence, and their locations are shown in Figure S2 (triangles for indels and stars for SNPs). Figure 4 shows the variation of the edulis allele along the sequence in the four patches. Whereas the frequency of the edulis allele remains consistently low in the Iberian Coast patch (Fqedu , 0.05) and consistently high in the North Sea and Bay of Biscay patches (Fqedu . 0.95), we noted a slight increase of the edulis allele toward the 59 side of mac-1 in the Brittany patch. However, there was a surpris- ing lack of variation along the whole sequence: the edulis allele frequency remains ^0.4 over several kilobase pairs.

Model fitting of adaptive introgression The limited scale of our chromosomal walking does not allow us to reject any hypotheses with high confidence. Accordingly, we next tested whether adaptive introgression might plausibly explain the pattern of diversity observed at mac-1.

First, to illustrate the effects of adaptive introgression on diversity, Figure 5 plots the expected reduction in diversity against the rate of gene flow (see also Pennings and Her- misson 2006). When M , 0.01 we see a clear "hard sweep" with a single haplotype reaching high frequency in the re- cipient deme, while when M . 1, most of the heterospecific diversity introgresses into the recipient background (a clas- sic soft sweep with no detectable reduction in diversity). For intermediate values of M, the effect is also intermediate, with both a soft sweep and a reduction in diversity.

Building on the results of Pennings and Hermisson (2006), we fitted a model of adaptive introgression to our mac-1 data. We first developed an ML approach to estimate the population mutation rate, u =2Nm, in the M. edulis populations (North Sea and Bay of Biscay), and the effective number of migrants, M =2Nm, from these populations to the Brittany patch of M. galloprovincialis (see Equation 2). Considering only edulis-derived alleles, the total number of alleles sampled, and the number of distinct allelic classes observed were nd = 177 and kd = 15 for the North Sea and Bay of Biscay patches and nr = 60 and kr = 6 for the Brittany patch (Table 3). From these data, we obtained a rel- atively precise estimate of the population mutation rate: ^u ¼ 3:79 [1.95 - 6.66] (see solid lines, Figure 6A). However, the effective number of migrants was imprecisely estimated, with low migration rates rejected but no upper bound: M^ ¼ 4:63 [0.95-] (see solid lines, Figure 6B). The lack of an upper bound on the estimate reflects the fact that nearly all of the allelic diversity from the edulis patches had been introgressed into the Brittany patch of M. galloprovincialis.

In an attempt to improve our inference, we next used an ABC approach, which allowed us to relax the assumption of strong selection and to consider more of the information in the data. Parameter posterior distributions, corrected for mean bias, are shown in Figure 6 as shaded circles. In the following, we give their median and 95% credible intervals. We obtained an even more precise estimate of the scaled mutation rate (Figure 6A): umedian = 2.82 [1.42 - 4.53] whose bounds are similar to the ML estimates. Estimation of the scaled migration rate (Figure 6B): Mmedian = 99 [24 - 154] showed a clear discrepancy between the two methods with respect to the lower bound. Compared to ML, the lower bound estimation was M ^ 1 while its upper bound was poorly estimated (value approximates the prior upper bound). Finally, the estimation of selective strength, amedian = 402 [203 - 594] (Figure 6C) was imprecise but compati- ble with the assumption of strong selection used in the ML inference.

To better understand the discrepancy between the two methods, we reran the ABC estimation, using only the summary statistics that appear in the ML equations (namely, nd, kd, nr, and kr). In this case, estimates of u and M were very similar to the MLEs (Figure S4), and there was no in- formation on a as expected. The results of the ABC approach therefore come from the additional use of summary statistics that contain information that is not used in the ML approach. We therefore conducted a goodness-of-fittesttoevaluate how well the model fit the data at mac-1 when estimating the parameters by each method (ML and ABC using all sta- tistics). Figure S5 shows the goodness-of-fit of ML estimation (blue dots) and ABC estimation (red dots) for each statistic. In the main, both methods fit the data well (see Table S1 for quantiles and P-values) but looking at the median values, ML performs better. In particular, ABC tends to overestimate kr and Her, which basically explains the discrepancy in the esti- mation of M. The ML method proved to be the most under- standable approach with a simpler model that retrieved the best fit to the data and consequently our interpretations rely on the ML estimates.

Finally, we used our simulations to assess the effects of violating some assumptions of our model: (1) the assumption of complete linkage between the selected and neutral loci and (2) the assumption that migration rates remained constant. Figure S6 shows that the sweep pattern was indeed robust to including recombination. Figure S7 suggests that our results will also be fairly robust to temporal variation in migration rates, particularly in mussels where dispersal rates are very high.

Insights into secondary contact history Both estimation methods agree that M . 1 best explains our data if adaptive introgression has taken place. To assess the plausibility of our estimation, we analyzed several models of speciation (Figure S8) between M. edulis and the Brittany patch of M. galloprovincialis.

We first applied a model choice procedure for each of the three speciation models with gene flow to test whether heterogeneity in introgression rates was supported. We observed clear support for heterogeneous migration (Table S3, "within model"). We then applied the model-choice pro- cedure to the speciation models, assuming heterogeneous migration (Table S3, "between models"). Secondary contact was the best-supported model with posterior probability, Ppost = 0.61, confirming a previous study (Roux et al. 2014). We tested the robustness of this result by simulating 500 pseudo-observed data sets for each model. We empirically estimated that given a posterior probability of 0.61, the probability that secondary contact with heterogeneous in- trogression rates is the correct model was 0.987 (Table S4). It is worth mentioning that when migration rate was assumed to be homogeneous among loci, the method failed to infer secondary contact; instead ancient migration was the best-supported model (SC, Ppost = 0.26 and AM, Ppost = 0.65).

We then estimated the multilocus distributions of in- trogression rates in the best-supported SC model with heterogeneous migration (Figure 7). We found an asymme- try of introgression in favor of M. edulis: Mgallo to edulis = 0.15 [0.002 - 1.06] (see open bars, Figure 7) and Medulis to gallo = 0.45 [0.013 - 2.84] (see shaded bars, Figure 7). However, compared to the multilocus inference, introgression at mac-1 was strongly unidirectional from M. edulis to M. gallopro- vincialis as suggested by the comparison of different scenar- ios of introgression (SC with introgression into M. edulis: Ppost = 0.0013; SC with introgression into M. galloprovincia- lis: Ppost = 0.9997; see Table S5). Furthermore introgression rate into M. galloprovincialis at mac-1 was very high com- pared to the extent of gene flow across loci (Mmedian = 2.50, solid line, Figure 7) that is consistent with the mac-1 gene- alogy (Figure 3). In comparison, our ML estimation of M at mac-1 in the adaptive introgression model (M^ ¼ 4:63, Fig- ure 6B) was outside the 95% quantile of the multilocus distribution, but its lower bound (MML lower bound = 0.95, dashed line, Figure 7) remained consistent with the second- ary contact inference.

Discussion Adaptive introgression between hybridizing species has long been proposed as a potentially important source of new adaptations in plants and, more recently, in animals too (Arnold 2004; Hedrick 2013). Most of all, compared to ad- aptation from new mutations or standing variation, adaptive introgression might allow big evolutionary jumps by the acquisition of complex variants (Wright 1949; Mallet 2007; Abbott et al. 2013).

Genome scan surveys of differential introgression among loci can help to localize genomic regions that introgress adaptively (Payseur 2010). Here, we took advantage of the mosaic structure of the hybrid zone of mussels along the western European coastline, to perform a multilocus scan of local introgression. As pointed out by Harrison and Rand (1989), an exciting feature of mosaic hybrid zones is that they involve independent contacts between species, each with a unique evolutionary trajectory. In addition to random fluctuations, the outcomes will be contingent to the relative abundance of the two species, local environmental condi- tions, and the genetic architecture of their reproductive isolation.

Our multilocus scan, based on AFLP and codominant markers, showed some genetic differences between the peripheral patch of M. galloprovincialis in the Iberian Coast and the patch enclosed in Brittany. In our analysis, AFLP markers were used mainly to obtain an estimation of the genomic distribution of FST values. We also have fitted a spe- ciation model to an independent DNA sequence data set. We therefore checked whether the parameters inferred could reproduce the FST distribution observed at AFLP markers. The two distributions-AFLPs vs. simulations under the best demographic model inferred-were very similar (Figure S9), suggesting that we have a good estimate of the genomic distribution of FST. To mitigate the problem of false posi- tives, due to an underestimation of the neutral variance of FST (Robertson 1975), we cross checked the results of six methods, all of which assume different population struc- tures. In this way, we identified three loci with a robust de- viation from the neutral expectations (open circles, Figure 2A). When compared to other methods, simulations of the inferred secondary contact model with heterogeneous gene flow produced an elevated variance of FST in both the between-species comparison, and-more surprisingly-in the within-species comparison. Nevertheless, even under this scenario, the combination of a high differentiation within and between species observed at mac-1 was not observed under neutrality (see dashed lines, Figure 2A). The geneal- ogy of mac-1 confirmed that this was due to local introgres- sion of M. edulis alleles into the Brittany patch of M. galloprovincialis, but not into the patch on the Iberian Coast (Figure 3). This pattern contrasts strongly with results at the other loci analyzed.

The locus-specific nature of the introgression at mac-1 suggested that selective processes were acting rather than purely demographic ones. The highly asymmetrical intro- gression rates from M. edulis (Table S5) contrasts with the multilocus pattern (Figure 7), and so we hypothesized that mac-1 might have been subject to adaptive introgression, with edulis alleles sweeping to fixation in the Brittany patch of M. galloprovincialis.

Demonstrating adaptive introgression requires a combi- nation of evidence that is hard to obtain in most species (Vekemans 2010). Few studies have undertaken experi- ments to confirm the fitness advantage of the introgressed traits/alleles in the recipient backgrounds. Exceptions in- clude the adaptive introgression of drought tolerance in Hel- ianthus annuus (Whitney et al. 2010) and the adaptive introgression of rodent poison resistance in the house mouse, Mus musculus domesticus (Song et al. 2011). In con- trast, most studies rely on indirect evidence involving geo- graphical variation in allele frequencies, analysis of DNA polymorphism, and the reconstruction of gene genealogies. Forexample,HeliconiusGenomeConsortium(2012),Pardo- Diaz et al. (2012), and Smith and Kronforst (2013) provided strong evidence for the adaptive introgression of wing color pattern genes involved in Müllerian mimicry between hy- bridizing species of Heliconius.Inhumans,Mendezet al. (2012a,b, 2013) found a high frequency of divergent hap- lotypes at two immune genes, STAT2 and OAS1, in modern Melanesians. These haplotypes share recent common ances- try with archaic hominin sequences (STAT2, Neanderthal; OAS1, Denisovan). This result is suggestive of ancient intro- gressions, but their adaptive nature remains hypothetical. Similarly Roux et al. (2013) identified genomic hotspots of introgression between two highly divergent sea squirt spe- cies but could draw no conclusions about their adaptive nature, illustrating that specific tests need to be developed.

The patterns that we observe at mac-1 differ in one re- spect from previously studied cases of outlying introgres- sion. In particular, in most previous cases, there has been a reduction in the genetic diversity in the recipient popula- tion, when compared to the donor population (see Table 1). In contrast, at the mac-1 locus, 96% of the M. edulis diversity was introgressed to M. galloprovincialis, and there were no other signs of a classical selective sweep (Table 1, Table 2, and Table 3).

However, it remains possible that the mac-1 alleles have introgressed via a soft sweep, with little loss of diversity (Pennings and Hermisson 2006). We tried two approaches to test for a soft sweep at this locus. First, we analyzed four polymorphisms along the sequence (Figure 4) with the hope of observing a gradient of introgression. Unfortunately, the chromosomal scale proved to be too small. Second, we used two new inference methods to test whether the rate of mi- gration required to generate the soft sweep effect was con- sistent with the existence of an efficient genetic barrier at mac-1 elsewhere in the range. Our inference suggested that M =2Nm . 1 was required to explain our data under the adaptive introgression model, and the ML estimate (M^ ¼ 4:63; Figure 6B) was outside of the 95% quantile of the multilocus distribution of introgression rates between M. edulis and the Brittany patch of M. galloprovincialis (Figure 7). However, its lower bound remained consistent with the secondary contact inference (MML lower bound = 0.95, dashed line, Figure 7), and so we cannot reject the hypothesis of adaptive introgression.

While we cannot reject adaptive introgression, we must also recognize alternative hypotheses. Migration pressure can easily spread gene combinations, even if these are universally selected against (e.g., Barton 1992). It is there- fore possible that our results could be explained by a geo- graphically and genomically localized change in the barrier strength between M. edulis and M. galloprovincialis. The genomic region around mac-1 may contain barrier loci that were maintained in some patches, but not in the Brittany patch-perhaps due to differences in local ecological selec- tion or because an asymmetric intrinsic incompatibility man- aged to uncouple from the tension zone.

Together, our results emphasize the difficulties in inter- preting outlier loci in populations that have undergone secondary contact. This is unsuprising given that patterns of introgression after such a contact can vary dramatically, depending on the way divergence occurred in allopatry, the resulting genetic architecture of the barrier, and the local landscape where the contact takes place-including its pop- ulation density, connectivity, and environmental variation (Bierne et al. 2011, 2013a; Domingues et al. 2012; Strasburg et al. 2012; Abbott et al. 2013; Gagnaire et al. 2013). Nev- ertheless, our best hope of understanding these issues is to apply dedicated methods, such as those developed here, to much larger genome-wide data sets.

Acknowledgments The authors are grateful to Marie-Thérèse Augé and Célia Gosset for their technical help with this work and Marc Ohresser who kindly provided the sequence of the mac-1 region. Molecular data were produced through the ISEM platform Génomique marine at the Station Méditerranenne de l'Environnement Littoral [OSU OREME (Observatoire de Recherche Méditerranéen de l'Environnement)] and the platform Génomique Environnementale of the LabEx CeMEB (Laboratoire d'Excellence Centre Méditerranéen de l'Environnement et de la Biodiversité). Numerical results presented in this article were carried out using the ISEM computing cluster at the platform Montpellier Bioinforma- tique et Biodiversité of the LabEx CeMEB, and we are grate- ful to its staff. This work was funded by the Agence National de la Recherche (HYSEA project, ANR-12-BSV7- 0011) and the project Aquagenet (SUDOE, INTERREG IV B). This is article 2014-044 of Institut des Sciences de l'Evolution de Montpellier (ISEM).

Literature Cited Abbott,R.,D.Albach,S.Ansell,J.W.Arntzen,S.J.E.Bairdet al., 2013 Hybridization and speciation. J. Evol. Biol. 26(2): 229-246.

Arnold, M. L., 2004 Transfer and origin of adaptations through natural hybridization: Were Anderson and Stebbins right? Plant Cell 16(3): 562-570.

Barton, N. H., 1979 The dynamics of hybrid zones. Heredity 43(3): 341-359.

Barton, N. H., 1992 On the spread of new gene combinations in the third phase of Wright's shifting-balance. Evolution 46(2): 551-557.

Barton, N. H., 2013 Does hybridization influence speciation? J. Evol. Biol. 26(2): 267-269.

Barton, N. H., and B. O. Bengtsson, 1986 The barrier to genetic exchange between hybridising populations. Heredity 57: 357-376.

Beaumont, M. A., and D. J. Balding, 2004 Identifying adaptive genetic divergence among populations from genome scans. Mol. Ecol. 13(4): 969-980.

Beaumont, M. A., and R. A. Nichols, 1996 Evaluating loci for use in the genetic analysis of population structure. Proc. Biol. Sci. 263(1377): 1619-1626.

Beaumont, M. A., W. Zhang, and D. J. Balding, 2002 Approximate Bayesian computation in population genetics. Genetics 162: 2025-2035.

Belkhir, K., P. Borsa, L. Chikhi, N. Raufaste, and F. Bonhomme, 1996-2004 GENETIX, version 4.05. http://kimura.univ-montp2. fr/genetix/ Bierne, N., 2010 The distinctive footprints of local hitchhiking in a varied environment and global hitchhiking in a subdivided population. Evolution 64(11): 3254-3272.

Bierne, N., P. David, P. Boudry, and F. Bonhomme, 2002 Assortative fertilization and selection atlarvalstageinthemusselsMytilus edulis and M. galloprovincialis. Evolution 56(2): 292-298.

Bierne, N., F. Bonhomme, and P. David, 2003a Habitat preference and the marine-speciation paradox. Proc. Biol. Sci. 270(1522): 1399-1406.

Bierne, N., P. Borsa, C. Daguin, D. Jollivet, F. Viard et al., 2003b Introgression patterns in the mosaic hybrid zone be- tween Mytilus edulis and M. galloprovincialis. Mol. Ecol. 12(2): 447-461.

Bierne, N., C. Daguin, F. Bonhomme, P. David, and P. Borsa, 2003c Direct selection on allozymes is not required to explain heterogeneity among marker loci across a Mytilus hybrid zone. Mol. Ecol. 12(9): 2505-2510.

Bierne, N., F. Bonhomme, P. Boudry, M. Szulkin, and P. David, 2006 Fitness landscapes support the dominance theory of post-zygotic isolation in the mussels Mytilus edulis and M. gallo- provincialis. Proc. Biol. Sci. 273(1591): 1253-1260.

Bierne, N., J. Welch, E. Loire, F. Bonhomme, and P. David, 2011 The coupling hypothesis: why genome scans may fail to map local adaptation genes. Mol. Ecol. 20(10): 2044-2072.

Bierne, N., P.-A. Gagnaire, and P. David, 2013a The geography of introgression in a patchy environment and the thorn in the side of ecological speciation. Curr. Zool. 59(1): 72-86.26 Bierne, N., D. Roze, and J. J. Welch, 2013b Pervasive selection or is it. . .?: Why are FST outliers sometimes so frequent? Mol. Ecol. 22(8): 2061-2064.

Blum, M. G. B., and O. Francois, 2010 Non-linear regression mod- els for approximate Bayesian computation. Stat. Comput. 20(1): 63-73.

Bonhomme, M., C. Chevalet, B. Servin, S. Boitard, J. Abdallah et al., 2010 Detecting selection in population trees: the Lewontin and Krakauer test extended. Genetics 186: 241-262.

Caballero, A., H. Quesada, and E. Roldan-Alvarez, 2008 Impact of amplified fragment length polymorphism size homoplasy on the estimation of population genetic diversity and the detection of selective loci. Genetics 179: 539-554.

Charalambides, C. A., and J. Singh, 1988 Review of the stirling numbers, their generalizations and statistical applications. Com- mun. Stat. Theory 17(8): 2507-2532.

Charlesworth, B., M. Nordborg, and D. Charlesworth, 1997 The effects of local selection, balanced polymorphism and back- ground selection on equilibrium patterns of genetic diversity in subdivided populations. Genet. Res. 70(2): 155-174.

Cornuet, J.-M., V. Ravigné, and A. Estoup, 2010 Inference on population history and model checking using DNA sequence and microsatellite data with the software DIYABC (v1.0). BMC Bioinformatics 11(1): 401.

Corpet, F., 1988 Multiple sequence alignment with hierarchical clustering. Nucleic Acids Res. 16(22): 10881-10890.

Csilléry, K., O. François, and M. G. B. Blum, 2012 abc: an R pack- age for approximate Bayesian computation (ABC). Methods Ecol. Evol. 3(3): 475-479.

Daguin, C., F. Bonhomme, and P. Borsa, 2001 The zone of sym- patry and hybridization of Mytilus edulis and M. galloprovincia- lis, as described by intron length polymorphism at locus mac-1. Heredity 86: 342-354.

Domingues, V. S., Y.-P. Poh, B. K. Peterson, P. S. Pennings, J. D. Jensen et al., 2012 Evidence of adaptation from ancestral var- iation in young populations of beach mice. Evolution 66(10): 3209-3223.

Edwards, A. W. F., 1992 Likelihood. Johns Hopkins University Press, Baltimore, MD.

Ewens, W. J., 1972 The sampling theory of selectively neutral alleles. Theor. Popul. Biol. 3(1): 87-112.

Excoffier, L., A. Estoup, and J.-M. Cornuet, 2005 Bayesian analysis of an admixture model with mutations and arbitrarily linked markers. Genetics 169: 1727-1738.

Fagundes, N. J. R., N. Ray, M. Beaumont, S. Neuenschwander, F. M. Salzano et al., 2007 Statistical evaluation of alternative mod- els of human evolution. Proc. Natl. Acad. Sci. USA 104(45): 17614-17619.

Faure, M. F., P. David, F. Bonhomme, and N. Bierne, 2008 Genetic hitchhiking in a subdivided population of Mytilus edulis.BMC Evol. Biol. 8: 164.

Fay, J. C., and C. I. Wu, 2000 Hitchhiking under positive Darwin- ian selection. Genetics 155: 1405-1413.

Gagnaire, P.-A., S. A. Pavey, E. Normandeau, and L. Bernatchez, 2013 The genetic architecture of reproductive isolation during speciation-with-gene-flow in lake whitefish species pairs as- sessed by RAD sequencing. Evolution 67(9): 2483-2497.

Gosset, C. C., and N. Bierne, 2012 Differential introgression from a sister species explains high FST outlier loci within a mussel species. J. Evol. Biol. 26(1): 14-26.

Hall, T., 1999 BioEdit: a user-friendly biological sequence align- ment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp. Ser. 41: 95-98.

Hamilton, G., M. Stoneking, and L. Excoffier, 2005 Molecular analysis reveals tighter social regulation of immigration in pat- rilocal populations than in matrilocal populations. Proc. Natl. Acad. Sci. USA 102(21): 7476-7480.

Harrison, R. G., 1986 Pattern and process in a narrow hybrid zone. Heredity 56(3): 337-349.

Harrison, R. G., and D. M. Rand, 1989 Mosaic hybrid zone and the nature of species boundaries, pp. 111-133 in Speciation and Its Consequences, edited by D. Otte, and J. A. Endler. Sinauer Associates, Sunderland, MA.

Hedrick, P. W., 2013 Adaptive introgression in animals: examples and comparison to new mutation and standing variation as sources of adaptive variation. Mol. Ecol. 22(18): 4606-4618.

Heliconius Genome Consortium, 2012 Butterfly genome reveals promiscuous exchange of mimicry adaptations among species. Nature 487(7405): 94-98.

Hewitt, G. M., 2011 Quaternary phylogeography: the roots of hybrid zones. Genetica 139(5): 617-638.

Hilbish, T. J., F. P. Lima, P. M. Brannock, E. K. Fly, R. L. Rognstad et al., 2012 Change and stasis in marine hybrid zones in re- sponse to climate warming. J. Biogeogr. 39(4): 676-687.

Hohenlohe, P. A., S. Bassham, P. D. Etter, N. Stiffler, E. A. Johnson et al., 2010 Population genomics of parallel adaptation in threespine stickleback using sequenced RAD tags. PLoS Genet. 6(2): e1000862.

Hudson, R. R., 2002 Generating samples under a Wright-Fisher neutral model of genetic variation. Bioinformatics 18(2): 337- 338.

Hudson, R. R., and N. L. Kaplan, 1985 Statistical properties of the number of recombination events in the history of a sample of DNA sequences. Genetics 111: 147-164.

Jackson, B., T. Kawakami, S. Cooper, J. Galindo, and R. Butlin, 2012 A genome scan and linkage disequilibrium analysis among chromosomal races of the Australian grasshopper Van- diemenella viatica. PLoS ONE 7(10): e47549.

Kemppainen, P., T. Lindskog, R. Butlin, and K. Johannesson, 2011 Intron sequences of arginine kinase in an intertidal snail suggest an ecotype-specific selective sweep and a gene duplica- tion. Heredity 106(5): 808-816.

Kumar, S., K. Tamura, and M. Nei, 2004 MEGA3: integrated soft- ware for molecular evolutionary genetics analysis and sequence alignment. Brief. Bioinform. 5(2): 150-163.

Lehman, N., A. Eisenhawer, K. Hansen, L. D. Mech, R. O. Peterson et al., 1991 Introgression of coyote mitochondrial DNA into sympatric North American gray wolf populations. Evolution 45(1): 104-119.

Lewontin, R. C., and J. Krakauer, 1973 Distribution of gene fre- quency as a test of the theory of the selective neutrality of poly- morphisms. Genetics 74: 175-195.

Mallet, J., 2007 Hybrid speciation. Nature 446(7133): 279-283.

Maynard Smith, J., and J. Haigh, 1974 The hitchhiking effect of a favourable gene. Genet. Res. 23(01): 23 - 35.

McDonald, J. H., 1994 Detecting natural selection by comparing geographic variation in protein and DNA polymorphisms, pp. 88-100 in Non-neutral Evolution, edited by B. Golding, Springer, New York.

McKinnon, G. E., R. E. Vaillancourt, D. A. Steane, and B. M. Potts, 2004 The rare silver gum, Eucalyptus cordata, is leaving its trace in the organellar gene pool of Eucalyptus globulus. Mol. Ecol. 13(12): 3751-3762.

Mendez, F. L., J. C. Watkins, and M. F. Hammer, 2012a Global genetic variation at OAS1 provides evidence of archaic admix- ture in melanesian populations. Mol. Biol. Evol. 29(6): 1513- 1520.

Mendez, F. L., J. C. Watkins, and M. F. Hammer, 2012b A haplo- type at STAT2 introgressed from neanderthals and serves as a candidate of positive selection in Papua New Guinea. Am. J. Hum. Genet. 91(2): 265-274.

Mendez, F. L., J. C. Watkins, and M. F. Hammer, 2013 Neandertal origin of genetic variation at the cluster of OAS immunity genes. Mol. Biol. Evol. 30(4): 798-801.

Nachman, M. W., and B. A. Payseur, 2012 Recombination rate variation and speciation: theoretical predictions and empirical results from rabbits and mice. Philos. Trans. R. Soc. B 367 (1587): 409-421.

Nadeau,N.J.,A.Whibley,R.T.Jones,J.W.Davey,K.K.Dasmahapatra et al., 2012 Genomic islands of divergence in hybridizing Heliconius butterflies identified by large-scale targeted sequencing. Philos. Trans. R. Soc. B 367(1587): 343-353.

Narita, S., M. Nomura, Y. Kato, and T. Fukatsu, 2006 Genetic structure of sibling butterfly species affected by Wolbachia in- fection sweep: evolutionary and biogeographical implications. Mol. Ecol. 15(4): 1095-1108.

Nei, M., 1973 Analysis of gene diversity in subdivided popula- tions. Proc. Natl. Acad. Sci. USA 70(12): 3321-3323.

Nei, M., 1978 Estimation of average heterozygosity and genetic dis- tance from a small number of individuals. Genetics 89: 583-590.

Pardo-Diaz, C., C. Salazar, S. W. Baxter, C. Merot, W. Figueiredo- Ready et al., 2012 Adaptive introgression across species bound- aries in Heliconius butterflies. PLoS Genet. 8(6): e1002752.

Parsons, T. J., S. L. Olson, and M. J. Braun, 1993 Unidirectional spread of secondary sexual plumage traits across an avian hy- brid zone. Science 260(5114): 1643-1646.

Payseur, B. A., 2010 Using differential introgression in hybrid zones to identify genomic regions involved in speciation. Mol. Ecol. Resour. 10(5): 806-820.

Pennings, P. S., and J. Hermisson, 2006 Soft sweeps II: molecular population genetics of adaptation from recurrent mutation or migration. Mol. Biol. Evol. 23(5): 1076-1084.

Pérez-Figueroa,A.,M.J.García-Pereira,M.Saura,E.Rolán- Alvarez,andA.Caballero,2010 Comparing three different methods to detect selective loci using dominant markers. J. Evol. Biol. 23(10): 2267-2276.

Piálek, J., and N. H. Barton, 1997 The spread of an advantageous allele across a barrier: the effects of random drift and selection against heterozygotes. Genetics 145: 493-504.

Quesada, H., R. Wenne, and D. O. Skibinski, 1999 Interspecies transfer of female mitochondrial DNA is coupled with role-reversals and departure from neutrality in the mussel Mytilus trossulus.Mol. Biol. Evol. 16(5): 655-665.

Rawson, P. D., and T. J. Hilbish, 1998 Asymmetric introgression of mitochondrial DNA among european populations of blue mussels (Mytilus spp.). Evolution 52(1): 100-108.

Robertson, A., 1975 Gene frequency distributions as a test of se- lective neutrality. Genetics 81: 775-785.

Roesti, M., D. Moser, and D. Berner, 2013 Recombination in the threespine stickleback genome-patterns and consequences. Mol. Ecol. 22(11): 3014-3027.

Ross-Ibarra,J.,S.I.Wright,J.P.Foxe,A.Kawabe,L.DeRose-Wilson et al., 2008 Patterns of polymorphism and demographic history in natural populations of Arabidopsis lyrata. PLoS ONE 3(6): e2411.

Roux, C., V. Castric, M. Pauwels, S. I. Wright, P. Saumitou-Laprade et al., 2011 Does speciation between Arabidopsis halleri and Arabidopsis lyrata coincide with major changes in a molecular target of adaptation? PLoS ONE 6(11): e26872.

Roux, C., G. Tsagkogeorga, N. Bierne, and N. Galtier, 2013 Crossing the species barrier: genomic hotspots of introgression between two highly divergent Ciona intestinalis species. Mol. Biol. Evol. 30(7): 1574-1587.

Roux, C., C. Fraisse, V. Castric, X. Vekemans, G. H. Pogson et al., 2014 Can we continue to neglect genomic variation in intro- gression rates when inferring the history of speciation?: a case study in a Mytilus hybrid zone. J. Evol. Biol. (in press).

Rozas, J., J. C. Sánchez-DelBarrio, X. Messeguer, and R. Rozas, 2003 DnaSP, DNA polymorphism analyses by the coalescent and other methods. Bioinformatics 19(18): 2496-2497.

Skibinski, D. O. F., J. A. Beardmore, and T. F. Cross, 1983 Aspects of the population genetics of Mytilus (Mytilidae; Mollusca) in the British Isles. Biol. J. Linn. Soc. Lond. 19(2): 137-183.

Slatkin, M., 1994 An exact test for neutrality based on the Ewens sampling distribution. Genet. Res. 64(1): 71-74.

Slatkin, M., 1996 A correction to the exact test based on the Ewens sampling distribution. Genet. Res. 68(3): 259-260.

Smith, J., and M. R. Kronforst, 2013 Do Heliconius butterfly spe- cies exchange mimicry alleles? Biol Lett. 9, 20130503.

Song, Y., S. Endepols, N. Klemann, D. Richter, F.-R. Matuschka et al., 2011 Adaptive introgression of anticoagulant rodent 33 poison resistance by hybridization between old world mice. Curr. Biol. 21(15): 1296-1301.

Strasburg, J. L., N. A. Sherman, K. M. Wright, L. C. Moyle, J. H. Willis et al., 2012 What can patterns of differentiation across plant genomes tell us about adaptation and speciation? Philos. Trans. R. Soc. B 367(1587): 364-373.

Tajima, F., 1983 Evolutionary relationship of DNA sequences in finite populations. Genetics 105: 437-460.

Tajima, F., 1989a The effect of change in population size on DNA polymorphism. Genetics 123: 597-601.

Tajima, F., 1989b Statistical method for testing the neutral muta- tion hypothesis by DNA polymorphism. Genetics 123: 585-595.

Tanguy, A., N. Bierne, C. Saavedra, B. Pina, E. Bachère et al., 2008 Increasing genomic information in bivalves through new EST collections in four species: development of new genetic markers for environmental studies and genome evolution. Gene 408(1-2): 27-36.

Turner, T. L., M. W. Hahn, and S. V. Nuzhdin, 2005 Genomic islands of speciation in Anopheles gambiae. PLoS Biol. 3: e285.

Vekemans, X., 2010 What's good for you may be good for me: evidence for adaptive introgression of multiple traits in wild sunflower. New Phytol. 187(1): 6-9.

Vekemans, X., T. Beauwens, M. Lemaire, and I. Roldán-Ruiz, 2002 Data from amplified fragment length polymorphism (AFLP) markers show indication of size homoplasy and of a re- lationship between degree of homoplasy and fragment size. Mol. Ecol. 11(1): 139-151.

Vitalis, R., K. Dawson, P. Boursot, and K. Belkhir, 2003 DetSel 1.0: a computer program to detect markers responding to selection. J. Hered. 94(5): 429-431.

Wakeley, J., and J. Hey, 1997 Estimating ancestral population parameters. Genetics 145: 847-855.

Watterson, G., 1975 On the number of segregating sites in genetical models without recombination. Theor. Popul. Biol. 7(2): 256-276.

Weir, B. S., and C. C. Cockerham, 1984 Estimating F-statistics for the analysis of population structure. Evolution 38(6): 1358- 1370.

Whitlock, R., H. Hipperson, M. Mannarelli, R. K. Butlin, and T. Burke, 2008 An objective, rapid and reproducible method for scoring AFLP peak-height data that minimizes genotyping error. Mol. Ecol. Res. 8(4): 725-735.

Whitney, K. D., R. A. Randell, and L. H. Rieseberg, 2010 Adaptive introgression of abiotic tolerance traits in the sunflower Helian- thus annuus. New Phytol. 187(1): 230-239.

Wright, S., 1949 Population structure in evolution. Proc. Am. Philos. Soc. 93(6): 471-478.

Zhivotovsky, L. A., 1999 Estimating population structure in dip- loids with multilocus dominant DNA markers. Mol. Ecol. 8(6): 907-913.

Communicating editor: M. A. Beaumont Christelle Fraïsse,*,[dagger],[double dagger],1 Camille Roux,*,[dagger],§ John J. Welch,[double dagger] and Nicolas Bierne*,[dagger] *Université Montpellier 2, 34095 Montpellier Cedex 5, France, [dagger]CNRS, Institut des Sciences de l'Evolution, ISEM Unité Mixte de Recherche 5554, 34200 SETE, France, [double dagger]Department of Genetics, University of Cambridge, CB2 3EH Cambridge, United Kingdom, and §Department of Ecology and Evolution, Lausanne University, Biophore/Sorge, CH-1015 Lausanne, Switzerland Copyright © 2014 by the Genetics Society of America doi: 10.1534/genetics.114.161380 Manuscript received January 10, 2014; accepted for publication April 17, 2014; published Early Online April 29, 2014.

Supporting information is available online at http://www.genetics.org/lookup/suppl/ doi:10.1534/genetics.114.161380/-/DC1.

1Corresponding author: SMEL, 2 rue des Chantiers, 34200 Sète, France. E-mail: [email protected] (c) 2014 Genetics Society of America

[ Back To TMCnet.com's Homepage ]