Abstract
Population genetic analysis of variation at five neutral microsatellite loci for Mediterranean mussels (Mytilus galloprovincialis) from 18 sites along the eastern Adriatic Sea revealed little or no spatial variation. In contrast, seascape genetics analysis revealed a pronounced locus‐specific gradient in allelic and genotypic frequencies across the study region. At a sixth locus, MGE7, the frequencies of two alleles, MGE7243 and MGE7249, were strongly associated, negatively and positively, respectively, with a single environmental variable – minimum salinity (minSAL). The frequency of the MGE7243/243 homozygous genotype was strongly negatively associated with minSAL, whereas the frequencies of the MGE7246/249 and the MGE7249/249 genotypes were strongly positively correlated with minSAL. Interpretation of these pronounced gradients is confounded by the fact that minSAL and another environmental variable, maximum sea surface temperature (maxSST), are highly correlated (R = −.911) and are therefore not necessarily acting independently. BLAST searches of the MGE7 locus against M. galloprovincialis whole genome shotgun sequence returned an alignment with contig mg10_S01094 (accession UYJE01010330.1) and 7 predicted M. galloprovincialis proteins VDI82194.1 ‐ VDI82200.1. Conserved domain searches revealed a similar structure to the transcriptional regulator Msx2‐interacting protein. The BLASTp search also returned significant alignments to Msx2‐interacting proteins in Mytilus coruscus, Crassostrea virginica, and Haliotis rubra. The existence of the MGE7 gradient highlights the role that environmental variation may play in retarding gene flow among wild M. galloprovincialis populations, and also how the success of collection of young mussels (spat) from one site and their transfer to another site (the farm) may be influenced by a single factor such as minSAL or maxSST on a localized scale.
Keywords: aquaculture; environmental variation; gene flow; genetic cline; Mediterranean Sea; Mollusca, Bivalvia; selection
We investigated the genetic variation of the economically and ecologically important species the Mediterranean mussel (Mytilus galloprovincialis) from 18 sites along the eastern Adriatic Sea, using six microsatellite loci. Population genetic analysis of variation at five neutral microsatellite loci revealed little or no spatial variation. In contrast, seascape genetic analysis revealed a pronounced association between allelic and genotypic frequencies at a sixth locus, MGE7, and the environmental variable minimum salinity, highlighting the role that environmental variation may play in retarding gene flow among wild M. galloprovincialis populations.

1. INTRODUCTION
Trying to understand the causative forces of patterns of population genetic variation at different spatial (and temporal) scales has long been a challenge for marine science (Schmidt et al., 2008; Selkoe et al., 2015). Classical population genetics is based on assumptions of marker neutrality because the investigator is usually trying to interpret gene flow (connectivity) in the absence of environmental variation – that is, without including the influence of selection because selective mortality may provide a picture of gene flow that is different from the actual (neutral) genetic connectivity (Holderegger et al., 2006). However, with the advent of highly polymorphic markers such as microsatellites, and more recently single nucleotide polymorphisms (SNPs), and with new analytical approaches, there is now potential to investigate how environmental variation may influence gene flow and population genetic structure (Gagnaire et al., 2015; Manel et al., 2003). The ability to generate new knowledge from markers that are themselves under selection (or are tightly linked to genes under selection) has implications for disciplines such as fisheries, aquaculture, and species conservation (Gaggiotti et al., 2009; Vaux et al., 2021; White et al., 2010), as well as providing new and broad insights into how organisms interact with their environments. For example, such new knowledge may be important in understanding species‐specific responses to environmental variation (e.g., increased sea surface temperature or changes in concentrations of aragonite/calcite and silica) in the face of climate change (e.g., Segovia et al., 2020; Wei et al., 2013a; Zeng et al., 2020).
For many benthic marine invertebrates, the main mechanism for exchange of individuals between populations is via a pelagic larval stage, the fate of which often depends on oceanographic features such as currents, fronts, straits, and shoreline configuration (Pascual et al., 2017; Riginos & Liggins, 2013). During the pelagic (dispersive) stage larvae may be transported within distinct packets of water, so environmental variables related to seawater may be relatively stable (Selkoe et al., 2006). With settlement in intertidal or shallow coastal regions where water mixing is often high, juveniles experience the sea moving about them bringing greater variation in environmental conditions (Riginos & Liggins, 2013). For sessile intertidal species, which cannot avoid the dynamic range of environmental stresses in the intertidal zone (Tomanek & Helmuth, 2002), the transition between pelagic and benthic stages may be a time of particular vulnerability (Pechenik, 1999). If extreme environmental conditions are a regular feature of the settlement location, selective mortality of juveniles can lead to genetic structuring in the adult populations. For example, peak spawning activity of Mediterranean mussels (Mytilus galloprovincialis Lamarck, 1819) in the Adriatic Sea occurs from late winter through to spring, a period that overlaps with times of pronounced freshwater influence (Da Ros et al., 1985; Krivokapić et al., 2011; Lipizer et al., 2014), a scenario that could lead to genetic structuring as a consequence of genotype‐dependent mortality in response to salinity fluctuation. This sort of environmentally driven selection has been very well documented in blue mussel (Mytilus edulis Linnaeus, 1758) populations in Long Island Sound (New York State, Atlantic coast of the USA), where regular annual selective mortality is driven by strong selection against juveniles with the Lap 94 allele in response to low salinity (Hilbish et al., 1982; Hilbish & Koehn, 1985).
A recent review of seascape genetics studies (Selkoe et al., 2016) noted that three factors – temperature, oceanography, and geography – were equally important in explaining patterns of spatial genetic variation across 100 studies. In addition, it was noted that many different factors are likely to affect connectivity at distinct spatio‐temporal scales, so while some studies may identify a single key (causative) environmental variable as explaining genetic variation, others will identify two or more variables, often acting synergistically (e.g., Silva & Gardner, 2016; White et al., 2010). It is increasingly clear that understanding patterns of spatial genetic variation, and the factors that contribute to the variation, is very much context‐ and species‐dependent. While important generalities may be drawn from reviews and meta‐analyses (e.g., Cárcamo, 2021; Selkoe et al., 2016), understanding the local situation is, of course, critical.
Our study focusses on M. galloprovincialis, a native species on the eastern shoreline of the Adriatic Sea. M. galloprovincialis is a model organism with a large body of research covering areas such as its role as an ecosystem engineer, physiological adaptations to the rocky intertidal zone, aquaculture and its potential as an invasive species (Arribas et al., 2014; Braby & Somero, 2006; Gardner et al., 2016; Kovačić et al., 2017; Wenne et al., 2022). Its success as an invader and ability to outcompete native congeners is in part due to its greater tolerance of a wide variety of environmental conditions in a time of global climate change (Branch & Steffani, 2004; Evans & Somero, 2010; Saarman et al., 2017; Tomanek, 2012). For example, on the Pacific coast of North America, it has supplanted the northern blue mussel (Mytilus trossulus Gould, 1850) in many locations from southern California (USA) to British Columbia (Canada) in waters where it is more warm‐tolerant than its native congener (Geller, 1999; Schneider & Helmuth, 2007; Tomanek, 2012).
The main circulation of the Adriatic Sea is cyclonic with the north‐flowing Eastern Adriatic Current (EAC) originating in the Ionian Sea and the Western Adriatic Current (WAC) flowing southward from the mouth of the Po River. Three cyclonic gyres, corresponding to northern, central, and southern basins, operate with varying strength through the year and significant variation between years may form barriers to larval dispersal and therefore lead to genetic discontinuities (Figure 1; Artegiani et al., 1997; Bray et al., 2017; Dubois et al., 2016; Zonn & Kostianoy, 2017). The Adriatic Sea is a dilution basin for the Mediterranean Sea, contributing ~30% of the freshwater inflow for the whole Mediterranean Sea (Estournel et al., 2021). The Adriatic Sea is nonetheless very saline with seasonal average salinity for the surface water of each of the three basins varying between 37.9–38.2, 37.7–38.3, and 35.5–37.4 psu for the southern, middle, and northern basins, respectively. Evaporation of water from the Adriatic Sea itself and the inflow of high‐salinity waters from the northern Ionian Sea (>38.6 psu) are the main processes that maintain this high salinity (Artegiani et al., 1997). The eastern Adriatic coastal region is characterized by a limestone karst landscape with accompanying erosion, which has led to an extensive and complex underground network of flow with many freshwater wells and springs, both on land and on the sea bottom (Bonacci et al., 2013). Mussels in these areas, as well as those in estuaries, can experience wide variations in salinity, which can change very quickly in times of heavy rain (UNEP/MAP‐RAC/SPA, 2015). Mussel tolerance to salinity variation may be as tied to the magnitude and speed of changes as to the absolute maximum and minimum values reached (Hamer et al., 2008).
FIGURE 1.

Sampling sites of mussels, M. galloprovincialis, showing (a) general currents: EAC eastern Adriatic current, WAC Western Adriatic current, NAdG northern Adriatic gyre, MAdG middle Adriatic gyre, SAdG southern Adriatic gyre (adapted from Lipizer et al., 2014), (b) the three regions used to test the hypothesis of genetic differentiation based on Adriatic Sea oceanography, (c) three sites in Mali Ston Bay, Bosnia & Herzegovina (NF) and Croatia (BF & BW), and (d) four sites in Boka Kotorska Bay (CW, CF, IF, IW), Montenegro. Site abbreviations as per Table 1.
The Mediterranean Sea is predicted to be particularly strongly influenced by global warming trends, with increasing occurrences of marine heat waves and temperature‐related mass mortality events (Di Camillo & Cerrano, 2015; Lejeusne et al., 2010; Michaelidis et al., 2014; Oliver et al., 2018). Conditions are already thought to be close to the upper limit of thermal tolerance for Mediterranean mussels in some regions with several instances of mass mortality associated with marine heatwaves in Spain, France, Greece, and in the Adriatic Sea (Galli et al., 2017; Lejeusne et al., 2010; Lupo et al., 2021; Michaelidis et al., 2014). Understanding existing levels of genetic diversity, patterns of gene flow, and the relationship between genetic variation, species ecology, physiology, and environmental factors is critical to the management of this ecologically and commercially important species. Understanding species biology in the native environment can also help in the management of the species in introduced habitats. Here, we describe the use of six polymorphic microsatellite markers to study the population and seascape genetics of 18 populations of M. galloprovincialis, along the eastern Adriatic coast.
2. MATERIALS AND METHODS
2.1. Sample collection
In total, 843 individuals of M. galloprovincialis were collected from 18 sites (hereafter populations) along the eastern coastline of the Adriatic Sea, a region spanning four countries (Croatia, Bosnia‐Herzegovina, Montenegro, and Albania) and a variety of coastal environments (Figure 1; Table 1). The 18 populations were divided between the three regions of the Adriatic Sea (defined by bathymetry and corresponding currents, see Figure 1) with six, seven, and five populations in the northern, central, and southern regions, respectively. Seven were farmed and 11 were wild populations. The number of individuals sampled per site varied between 31 and 50 (mean ± SD of 46.8 ± 5.6). A variety of different sizes (shell length) and therefore putative ages of mussels was collected per site, with shell length in the range of 8–85 mm. We collected all wild (native) mussels from the intertidal zone and the farmed mussels were collected from the shallow subtidal region (top 2 m of the water column at the farm). Samples of mantle tissue were preserved in 99% ethanol for DNA extraction.
TABLE 1.
Mytilus galloprovincialis sampling site code and name, status (W, Wild; F, Farmed), sample size (N), collection date, and geospatial coordinates.
| Site code | Site name | Wild (W) farmed (F) | Country | N | Collection date | Latitude | Longitude |
|---|---|---|---|---|---|---|---|
| LKF | Limski Channel | F | Croatia | 31 | 09/2015 | 45.13334 | 13.66667 |
| CK | Črišnjevo Krk bridge | W | Croatia | 43 | 09/2015 | 45.24935 | 14.58243 |
| KLW | Klimno, Soline Bay | W | Croatia | 49 | 09/2015 | 45.15497 | 14.61889 |
| OSOR | Cres‐Lošinj bridge | W | Croatia | 35 | 09/2015 | 44.69283 | 14.39231 |
| ML | Mali Lošinj Harbor | W | Croatia | 50 | 09/2015 | 44.53228 | 14.46835 |
| MLC | Mali Lošinj Čikat Bay | W | Croatia | 50 | 09/2015 | 44.53181 | 14.45143 |
| Z | Gaženica Harbor, Zadar | W | Croatia | 50 | 09/2015 | 44.09786 | 15.25289 |
| U | Ugljan‐Pašman bridge | W | Croatia | 50 | 09/2015 | 44.01455 | 15.25086 |
| STF | Poljica, near Split | F | Croatia | 48 | 09/2015 | 43.51543 | 16.13917 |
| STW | Vranjic, Split | W | Croatia | 48 | 09/2015 | 43.53180 | 16.46667 |
| NF | Neum, Mali Ston Bay | F | Bosnia & Herzegovina | 50 | 09/2015 | 42.91312 | 17.62381 |
| BF | Bistrina, Mali Ston Bay | F | Croatia | 50 | 09/2015 | 42.86880 | 17.69835 |
| BW | Bistrina, Mali Ston Bay | W | Croatia | 50 | 09/2015 | 42.87976 | 17.70917 |
| CF | Boka Kotorska Bay | F | Montenegro | 50 | 09/2015 | 42.48539 | 18.74596 |
| CW | Boka Kotorska Bay | W | Montenegro | 50 | 09/2015 | 42.48568 | 18.74349 |
| IF | Boka Kotorska Bay | F | Montenegro | 50 | 09/2015 | 42.43705 | 18.76321 |
| IW | Boka Kotorska Bay | W | Montenegro | 44 | 09/2015 | 42.43596 | 18.76367 |
| SBF | Butrint Lagoon | F | Albania | 45 | 05/2017 | 39.75474 | 20.03251 |
| Total | 843 |
2.2. Laboratory protocols
DNA was extracted using Geneaid genomic DNA Mini Kits following manufacturer's instructions and DNA concentrations were quantified using a NanoDrop™ ND‐1000 (Thermo Scientific). Individuals were genotyped using seven microsatellite loci described in three published papers as follows: Mgμ1 (Presa et al., 2002), MGE3, MGE4, MGE5, MGE7 (Yu & Li, 2007), MT203, MT282 (Gardeström et al., 2008). The choice of microsatellites was based on amplification success and the presence of polymorphism; details, including amplification conditions that followed Westfall (2011), are summarized in Table A1. Amplified fragments were scanned using an automated analyzer ABI 3730XL by two different service providers (Macrogen and Sangon Biotech).
2.3. Genetic data sets
Electropherograms were scored with Geneious 8.0.5 (http://www.geneious.com, Kearse et al., 2012) and Microchecker (Van Oosterhout et al., 2004) was used to check the microsatellite data for genotyping artifacts such as null alleles, large allele dropout, and stuttering. An outlier analysis to detect loci under selection was performed with Arlequin 3.5.2.2 (Excoffier & Lischer, 2010) using the Finite Island Model, 100 simulated demes and 20,000 coalescent simulations.
Different versions of the data set have been used for analyses (Table A2). This is because locus Mgμ1 had data missing for ~56% of all mussels and because locus MGE7 was potentially under selection (see outlier analysis section in Results). We used seven loci for all analyses, but repeated analyses with six loci (excluding Mgμ1). We present results for six loci only because results using seven loci and six loci (with and without Mgμ1) were similar. For the AMOVA, Structure and AWclust analyses, we also report results for five loci (i.e., excluding Mgμ1 and MGE7) so that results using neutral loci only may be compared directly to results that included MGE7. To calculate the F‐statistics used in the seascape analyses, we used a data set of 577 individuals containing complete data for six loci (i.e., excluding Mgμ1 because of missing data).
To assess whether sample sizes were sufficiently large to characterize the allelic variation of each population, the rarefaction procedure in the software package PopGenKit (Rioux Paquette, 2011) in R 3.5.1 (R Core Team, 2014) was used with 1000 jackknife replicates to estimate the total allelic diversity for each marker, each population and over all populations.
Analyses of Hardy–Weinberg equilibrium (HWE) and Linkage Disequilibrium (LD) were performed using Genepop 4.7.0 (Rousset, 2008). Departures from HWE were assessed by two methods, the probability test and the U‐test with the alternate hypothesis of heterozygote deficiency, using the default parameters. False discovery rate control for multiple statistical testing was applied to p‐values (Verhoeven et al., 2005).
The effective population size, Ne, was estimated using the linkage disequilibrium method of Hill (1981), Waples (2006), and Waples and Do (2010) as implemented in NeEstimator 2.1 (Do et al., 2014). Rare alleles were excluded from the calculation as recommended by Waples and Do (2010).
2.4. Allelic and genotypic estimations and analyses
The frequency of alleles, number of private alleles (PA), observed and expected heterozygosities (HO and HE), and the fixation index (F IS) were calculated using GenAlEx 6.503 (Peakall & Smouse, 2012). Allelic richness (AR) and private allelic richness (PAR) were calculated using HP‐RARE 1.1 (Kalinowski, 2005). Kruskal–Wallis tests were employed to test whether F IS and AR from farmed and wild populations were drawn from the same distribution. To test the hypothesis that alleles and genotypes are drawn from the same distribution in the 18 populations, exact G‐tests for population differentiation were performed in Genepop 4.7.0 for all populations using the default Markov chain parameters. The correlation between latitude and the frequency of the most common MGE7 allele was tested.
2.5. Differentiation among regions and populations
To test if there were genetic differences among populations in the three regions in the Adriatic Sea (Figure 1), a hierarchical analysis of molecular variance (AMOVA, Excoffier et al., 1992) was performed using GenAlEx 6.503 (Peakall & Smouse, 2012) for three regions and 18 populations with 999 permutations to determine statistical significance of the F RT values. Within AMOVA, the F‐statistics were calculated for individual loci to find the contribution of each locus to the overall differentiation indices. The p‐values were corrected for family‐wise error using the False Discovery Rate formula. Pairwise population F ST values were calculated to measure genetic differentiation between populations.
Two different cluster analysis programs, Structure (Pritchard et al., 2000) and AWclust (Gao & Starmer, 2008), were used to assess genetic structure for the 18 populations. Structure uses Bayesian analysis, which assumes conformation to HWE and LD assumptions, while AWclust is non‐parametric and therefore does not require conformity to HWE and LD assumptions. Structure analyses were performed with a burn‐in length of 50,000 steps, run length of 100,000 steps, 10 iterations for K = 1 to 18 using the correlated allele frequency model, and both the Admixture and No admixture models with sampling locations as priors. The Evanno method (Evanno et al., 2005) as implemented in the program Structure Harvester (Earl & vonHoldt, 2012) was used to detect the most appropriate number of clusters. AWclust analysis followed the allelic coding protocol (Gruber et al., 2013; Wei et al., 2013b) of 0 for no allele, 1 for heterozygous, and 2 for homozygous variants. To determine the most appropriate number of clusters, K, the gap statistic was calculated for K = 1–8 using 100 null simulations. For both Structure and AWclust, linear correlations were performed between latitude and the proportion of cluster membership for each population to see if cluster membership followed a latitudinal cline similar to MGE7 allele frequency.
2.6. Environmental data acquisition
Environmental data were collected from published articles and online databases (Table A3). Because there is no standardized environmental data collection between countries bordering the eastern Adriatic Sea, the data available varied widely among sites. Data from published articles were more plentiful for farmed and neighboring wild populations (LKF, NF, BF, BW, CF, CW, IF, IW, and SBF) and populations in large harbors (Z, STW, IF, IW) than for the other sites (CK, KLW, OSOR, ML, MLC, U, and STF) (see site abbreviations in Table 1). For these latter areas, which were all in Croatia, the online resources at SeaDataNet.org and the Croatian sea bathing water quality database (IOR, 2022), together with the models presented in Lipizer et al. (2014) and satellite data presented in Böhm et al. (2003), were used and also allowed for comparison among all sites. We did not use satellite data available from Copernicus because we could not retrieve data from enclosed bays and convoluted coastlines that accurately reflected fine‐scale conditions (Đurović et al., 2018). More detailed information about environmental data types, spatial and temporal coverage, and sources is provided in Hamilton (2019). Overall, there was good agreement for trends and inter‐site comparisons across all data sources, particularly for sea surface temperature and salinity data. The chlorophyll‐a data were not so complete, with detailed information available for the 10 southernmost populations but only rather sparse data for the eight northern populations, for which we relied on satellite‐derived data from Böhm et al. (2003) and the website SeaDataNet.org.
Nine environmental variables were used in analyses, the minimum and maximum values of sea surface temperature, salinity, and chlorophyll‐a concentrations (minSST, maxSST, minSAL, maxSAL, minCHL‐a and maxCHL‐a), together with delta values (deltaSST, deltaSAL and deltaCHL‐a – the difference between the maximum and minimum readings) that reflect the full range of variability. Three geospatial variables were also employed, latitude (Lat), longitude (Long), and the total coastal distance (TotalCD). TotalCD was calculated following Wei et al. (2013a) as the total of the pairwise coastal distances (CD in km) between any one population and all others (where CD is the shortest distance by sea between each population pair). Sites with high values of TotalCD are, on average, at a greater distance from all others than sites with low values of TotalCD. The environmental and geospatial variables used are detailed in Table A4.
2.7. Correlation testing of environmental/geospatial variables
To test for independence correlations between pairs of all twelve variables (nine environmental and three geospatial) were calculated and relationships visualized using principal component analysis using Statistica 7.1 (StatSoft). We removed variables so that |R| < 0.85 for all pairs of variables to avoid excess collinearity (e.g., Anderson et al., 2008).
2.8. Seascape genetics ‐ GLZ analysis
A generalized linear model (GLZ in Statistica) was used to describe the relationship between a genetic response variable and site‐specific environmental variation. GLZ analysis was carried out for five separate metrics: (1) LinF ST5 (i.e., MGE3, MGE4, MGE5, MT203, MT282) which is based on the complete data set of five neutral loci, (2) LinF ST6 (i.e., MGE3, MGE4, MGE5, MT203, MT282 plus MGE7) which includes the locus that is putatively under selection, (3) LinF STMGE7 which was based only on the MGE7 locus only, (4) f(MGE7243) ‐ the site‐specific frequency of allele MGE7243, and (5) f(MGE7243/243) ‐ the site‐specific frequency of the homozygous genotype. Estimates of linearized F ST (LinF ST) were calculated using GenAlEx 6.503 (Peakall & Smouse, 2012) following Wei et al. (2013a). These metrics focus progressively on the MGE7 locus and the 243 bp allele, starting with a measure containing no information from the MGE7 locus and finishing with measures focused on the frequency of the 243 bp allele and homozygote. GLZ analyses were carried out using six predictor (independent) variables: minSAL, minSST, maxSAL, minCHL‐a, Lat, and TotalCD.
GLZ with a normal distribution and a Log‐link function was employed to identify the best‐fit model based on the Akaike information criterion (AIC). For each of the five genetic indices, the top‐ranked model (lowest AIC value) and a set of best‐fitting models were used to evaluate which of the individual or combined predictor variables were most important in explaining variation in the genetic index. The number of models selected for the set of best‐fitting models was chosen by inspecting the plot of AIC values against the number of models and finding a point of inflection in the plot. This inflection is the point at which successive models add only a small and relatively constant amount to the AIC and are therefore not additionally informative (this is equivalent to the inspection of a scree plot for a PCA). The importance of each independent variable in explaining genetic structure was evaluated by calculating the percentage of best‐fitting models containing that independent variable (Figure A1; Table A5).
2.9. Seascape genetics ‐ distance‐based linear modeling
To test the effect of the six independent environmental/geospatial variables (minSAL, minSST, maxSAL, minCHL‐a, Lat, TotalCD) on site‐specific genetic variation, a distance‐based linear model was employed (e.g., Silva & Gardner, 2016). The DISTLM routine (PRIMER v6 with PERMANOVA add‐on ‐ Anderson et al., 2008) was used to test for associations between genetic and environmental variation (9999 permutations). The environmental and geospatial data were normalized before use. Genetic data were not transformed (i.e., raw population‐specific allele frequency data were employed) and a Euclidean distance similarity matrix was generated. Because this is a permutational test there are no assumptions about data normality (Anderson et al., 2008). The contribution of individual variables to explaining variation in the genetic data set was tested using marginal tests. Then, a model building approach was employed that tests all possible models and that identifies the best‐fit model containing successively more variables based on the Akaike Information Criterion (AIC), while also providing information about greatest improvement of the correlation coefficient (R 2). DISTLM analysis was carried out on the 6‐locus data set (i.e., excluding Mgμ1).
2.10. Correlation of MGE7 alleles and genotypes with environmental variables
We tested MGE7 alleles and genotypes for correlation with environmental variables that were important in explaining population‐specific genetic variation in the GLZ and DISTLM analyses. We tested correlations for all alleles and genotypes with frequencies greater than 0.05 (four alleles ‐ MGE7240, MGE7243, MGE7246, MGE7249 and five genotypes ‐ MGE7243/243, MGE7243/246, MGE7246/246, MGE7246/249, MGE7249/249) and also for the heterozygote genotype MGE7243/X where X is any other allele.
3. RESULTS
3.1. Genetic data and use of different data sets
Six of seven microsatellite loci were amplified satisfactorily with 85.1%–98.7% of individuals scored. Locus Mgμ1 showed stuttering and multiple peaks for a proportion of samples, with 44.1% of individuals able to be reliably scored overall and the percentage scored per population ranging between 28.0% and 64.7%. Because results of analyses with and without this unreliable locus were similar, we have only reported results not including this locus. There was no relationship between percentage scored per population and population latitude (R 2 = .0004, p = .94). The percentage of each locus scored per population was broadly consistent between the two service providers.
The outlier analysis did not detect any significant deviations in the F ST vs. heterozygosity among the loci. The high F ST value of the MGE7 locus, together with results from correlation testing, AMOVA, Structure, and AWClust, which clearly show MGE7 behaving differently from the other loci, did however suggest that MGE7 might in fact be an outlier that was not identified statistically because of the low number of loci used. We interpret this conservatively and suggest that this locus may be under selection or is located close to a gene under selection. As a consequence, we present analyses for the neutral (five loci, excluding MGE7) and the six locus data set (including MGE7).
The application of Microchecker identified an excess of homozygote genotypes across all 18 populations for four of the six loci (MT282, MGE4, MGE7, and MT203). For the remaining two loci, MGE3 and MGE5, only one and three populations, respectively, had excess homozygotes. All 18 populations showed significant departures from HWE expectations with four individual loci being out of HWE at every population. Loci MGE3 and MGE5 met expectations at 15 and 11 populations, respectively. There was no evidence of linkage disequilibrium for any pair of loci, either at population level or across all populations.
Rarefaction curves showed that ~80% of allelic variation was captured in each population for six loci estimates of the effective population size, Ne, were large (Table A6). For the majority of populations, the upper limit of the 95% confidence interval was “infinite,” where an “infinite” population size means that there is no evidence of genetic variation caused by genetic drift due to a finite number of parents. Only one population, IW, had a finite upper limit for the 95% confidence interval for both data sets.
3.2. Allelic and genotypic estimations and analyses
Average allelic richness, AR, ranged from 6.16 in NF to 7.90 in CK, and private allelic richness, PAR, ranged from 0.00 in SBF to 0.40 in STF. In total, there were 15 private alleles. Observed heterozygosity, H O, expected heterozygosity, H E, and the inbreeding coefficient, F IS, was similar for the six and seven locus data sets. H O was lower than H E at every population, with H O = 0.389 and H E = 0.629 across all populations (using six loci). This is reflected in the high inbreeding coefficient, F IS, for which values ranged from 0.267 to 0.453 across all populations and was 0.365 overall (Table A7). The individual F IS for each locus averaged over all populations varied considerably. Loci with low levels of homozygote excess (MGE3 and MGE5) had lower F IS , 0.052 and 0.115, respectively, compared to the loci that did have excess homozygotes at every population (MT282, 0.583; MGE4, 0.483; MGE7, 0.601; and MT203, 0.365).
The probability that alleles from each population were drawn from the same distribution for six loci, calculated by exact G‐tests, was <1 × 10−20 across all loci, less than 1 × 10−5 for individual loci MT282, MGE4, MGE7, and MT203, 0.021 for MGE5 and 0.187 for MGE3. The probability that genotypes from each population were drawn from the same distribution for six loci was 4.4 × 10−7 across all loci, less than 1 × 10−5 for MGE7, 0.018 for MT282 and not significant for the remaining loci.
There were no differences in AR or F IS between farmed and wild populations (Kruskal‐Wallis H statistic = 0.740, p = .39 and H statistic = 2.628, p = .10 for AR and F IS, respectively). Regional allele frequencies are illustrated in Figure A2. The frequency of the most common allele of the MGE7 locus, 243 bp in length, MGE7243, showed an increasing southward trend, with average frequency 46.5% in the North (range 39.1%–54.4%), 51.4% in the Central region (range 45.8%–59.5%), and 66.9% in the South (range 54.9%–77.0%) (Figure 2). The frequency of MGE7243 at each population was negatively correlated with latitude (R 2 = .476, p = .0015) (Figure 3).
FIGURE 2.

Allele frequencies for MGE7 for each population, each region and across all data.
FIGURE 3.

MGE7243 allele frequency and structure results: (a) map of frequency of MGE7243 allele (cyan) and all other alleles (orange) at each population, (b) correlation of frequency of MGE7243 allele with latitude, (c) map of proportion of cluster membership for six loci from the structure analysis (cluster 1 orange and cluster 2 cyan), (d) correlation of proportion of cluster membership by population for six loci (green) and five neutral loci (blue).
3.3. Differentiation among regions and populations
AMOVA (Table 2) revealed small but significant genetic differentiation among the regions for all six loci (F RT = 0.0031, p < .001) but not for the five neutral loci (F RT = 0.0007, p = .085). The F ST index, which in the hierarchical AMOVA is an index of variation explained between regions and populations to total variation, was significant for all six loci (F ST = 0.008, p < .001) and also for the five neutral loci (F ST = 0.002, p = .024).
TABLE 2.
Hierarchical AMOVA and differentiation indices using six loci and five neutral loci, for three regions and 18 populations
| Source of variation | Six loci | Five neutral loci | Index | |||||||
|---|---|---|---|---|---|---|---|---|---|---|
| Df | SS | Est var | % var | Index | Df | SS | Est var | % var | ||
| Among regions | 2 | 14.7 | 0.006 | 0.31% |
F RT = 0.0031 p < .001 |
2 | 6.3 | 0.001 | 0.07% |
F RT = 0.0007 p = .085 |
| Among populations | 15 | 58.4 | 0.010 | 0.52% |
F SR = 0.0052 p < .001 |
15 | 37.8 | 0.002 | 0.12% |
F SR = 0.001 p = .108 |
| Among individuals | 817 | 2387.9 | 0.909 | 44.7% |
F ST = 0.0082 p < .001 |
817 | 1850.4 | 0.665 | 39.69% |
F ST = 0.002 p = .024 |
| Within individuals | 835 | 923.0 | 1.105 | 54.4% |
F IS = 0.451 p < .001 |
835 | 815.5 | 1.007 | 60.12% |
F IS = 0.398 p < .001 |
| 2.031 | 100.0% |
F IT = 0.456 p < .001 |
1.675 | 100.00% |
F IT = 0.398 p < .001 |
|||||
Abbreviations: % var, percentage variation; Df, degrees of freedom; Est var, estimated variation; SS, sum of squares.
For the six locus data set, all three pairwise comparisons of F RT values among regions were statistically significant (p ≤ .012) whereas for the five locus data set only the North‐Central comparison was significant (Table 3).
TABLE 3.
Pairwise F RT for comparisons between the three regions are below the diagonal with p‐values following 9999 permutations above the diagonal
| Six loci | Five loci | ||||||
|---|---|---|---|---|---|---|---|
| Region | North | Central | South | Region | North | Central | South |
| North | 0.012 | ≤0.0001 | North | 0.013 | 0.146 | ||
| Central | 0.001 | ≤0.0001 | Central | 0.002 | 0.376 | ||
| South | 0.008 | 0.004 | South | 0.001 | 0.000 | ||
Note: Results for all six loci are on the left and for the five neutral loci on the right.
Pairwise population F ST values for the six locus and five neutral locus data sets were variable with 64 of 153 comparisons significant after FDR correction for the six locus data set and only one for the neutral data set (Table A8). Two populations exhibited unusually high F ST values for the neutral data set, ML and Z, both with mean an F ST value of 0.008 compared to mean values of 0.001 to 0.004 for all other population pairwise comparisons.
The Structure analysis runs using the Admixture model for the six locus data set had a maximum likelihood statistic, ΔK, at K = 5, with a local maximum at K = 2. When the No Admixture model was used, the maximum likelihood was attained at K = 2 (Figure A3). The Structure runs for K = 5 did not show any obvious genetic structure but for K = 2, using both models, the bar plot suggested a north–south cline in the proportion of ancestry from each cluster, something that was confirmed by linear correlation tests of proportion of cluster membership as a function of latitude (R 2 = .459, p = .002, Figures 3 and A3). Structure runs on the neutral data set did not show this north–south cline (R 2 = .111, p = .178, Figure 3) and there was no evidence of any genetic differentiation.
AWclust analyses (Figure A4) of the six locus data set revealed overlapping confidence intervals and local maxima at K = 2, 4, and 7, indicating weakly defined structure. For K = 2, the correlation between cluster 2 membership and latitude was significant (R 2 = .308, p = .017) but weaker than the corresponding correlation from the Structure program. For the five neutral loci, only one cluster was evident and when K = 2, the correlation between cluster 2 membership and latitude was not significant (R 2 = .073, p = .279).
3.4. Correlation testing of environmental/geospatial variables
Based on PCA and correlation analysis with a critical cut‐off value of |R| > 0.85, we removed six of 12 environmental and geospatial variables from the data set (Figure A5, Table A9). Subsequent analyses involved six independent environmental and geospatial variables (minSAL, minSST, maxSAL, minCHL‐a, Lat, and TotalCD).
3.5. Seascape genetics ‐ GLZ analysis
GLZ analyses were performed on five genetic response indices ‐ LinF ST5, LinF ST6, LinF STMGE7, f(MGE7243), and f(MGE7243/243) (Table 4). The top‐ranked models became more significant statistically as the genetic response index honed in on the MGE7243 allele: p = .177 for neutral loci (LinF ST5), p = .071 for all loci including MGE7 (LinF ST6), p = .023 for MGE7 only (LinF STMGE7), p < .0001 for both allele frequency (f(MGE7243)) and genotype frequency (f(MGE7243/243)). The variable minSAL was the most important of all predictor variables in explaining genetic variation: it was included in the top‐ranked model for all genetic indices except LinF ST5 and had a significant all effects test for all genetic indices except LinF STMGE7. It was included in an increasing percentage of the best‐fitting models as the genetic index focused on the MGE7243 allele (it was included in 12.5% of best‐fitting models for LinF ST5, 70% for LinF ST6, 73.3% for LinF STMGE7, 100% for both f(MGE7243) and f(MGE7243/243)). Although minSST and minCHL‐a were included in the top‐ranked model (together with minSAL) for f(MGE7243/243), the three lowest ranked models of all possible models included only minSST and/or minCHL‐a.
TABLE 4.
Results from GLZ analysis showing the top‐ranked models for variation in five genetic indices, LinF ST5, LinF ST6, LinF STMGE7, f(MGE7243), and f(MGE7243/243).
| Genetic index | Top‐ranked model | Best‐fitting models | Intercept p | ||||||
|---|---|---|---|---|---|---|---|---|---|
| Model | p | minSAL | minSST | maxSAL | minCHL‐a | Lat | TotalCD | ||
| LinF ST5 neutral loci | minSST | .177 | 12.5 | 50 | 25 | 0 | 25 | 25 | .014 |
| LinF ST6 six loci | minSAL Lat | .071 | 70 | 30 | 30 | 10 | 80 | 20 | .010 |
| LinF ST MGE7 | minSAL Lat | .023 | 73.3 | 53.3 | 53.3 | 13.3 | 73.3 | 26.7 | .143 |
| f(MGE7243) | minSAL | <.0001 | 100 | 28.6 | 28.6 | 14.331.6 | 14.3 | 14.3 | .400 |
| f(MGE7243/243) | minSAL minSST minCHL‐a | <.0001 | 100 | 80 | 40 | 46.7 | 33.3 | 26.7 | .108 |
Note: The top‐ranked model for each test is shown with p‐value, together with the percentage of each variable included in the set of best‐fitting models: Significant values according to the all‐effects test shown in bold. The p‐values for the all‐effects test on the intercept (null hypothesis: The intercept is at zero) are also shown.
3.6. Seascape genetics ‐ distance‐based linear modeling
Marginal analyses revealed that two of the six variables explained significant variation in the 6‐locus data set (minSAL (p = .0016), Lat (p = .0434)) (Table 5). Model building based on AIC values revealed that the single best‐fit model contained only one variable – minSAL (Table 5). This model explained 13.8% of the variation in the genetic data set. As expected, addition of new terms to the model improved the fit as judged by the R 2 value, which increased from 0.138 (one variable) to 0.432 (six variables). The variable minSAL was included in all six of the six best‐fit models, indicating that it is the single variable with greatest power to explain variation in the genetic data set.
TABLE 5.
Results of the distance‐based linear modeling (DISTLM) analyses, including marginal tests (per individual variable) and best‐fit models for all possible number combinations of variables.
| Marginal tests | |||
|---|---|---|---|
| Variable | Pseudo‐F | p | Proportion |
| minSAL | 2.560 | .0016 | 0.138 |
| Lat | 1.776 | .0434 | 0.100 |
| maxSAL | 1.490 | .1035 | 0.085 |
| minCHL‐a | 1.067 | .3372 | 0.062 |
| minSST | 0.933 | .5293 | 0.055 |
| TotalCD | 0.762 | .6624 | 0.045 |
| Best models ‐ best result for each number of variables | |||
|---|---|---|---|
| No. of variables | AIC | R 2 | Variables selected to be in the model in model |
| 1 | −38.543 | .138 | minSAL |
| 2 | −37.630 | .188 | minSAL + minCHL‐a |
| 3 | −37.434 | .266 | minSAL + minSST + maxSAL |
| 4 | −36.893 | .323 | minSAL + minSST + maxSAL + Lat |
| 5 | −36.768 | .390 | minSAL + minSST + maxSAL + Lat + TotalCD |
| 6 | −36.064 | .432 | minSAL + minSST + maxSAL + Lat + TotalCD + minCHL‐a |
3.7. Correlation of MGE7 alleles and genotypes with environmental variables
The most common allele over all populations, MGE7243, was negatively correlated with minSAL (f = 0.538, R 2 = .635, p < .0001). This was partially counterbalanced by the positive correlation of MGE7249 (f = 0.124, R 2 = .432, p = .003). The second most common allele, MGE7246, was not correlated with minSAL (f = 0.263, R 2 < .001, p = NS) (Figure 4).
FIGURE 4.

(a) Correlation of MGE7 allelic frequencies with minSAL, for alleles with overall frequency >0.05, showing frequency (f) of each allele over all populations, the correlation coefficient (R 2), gradient, and p‐value. (b) Correlation of MGE7 genotypic frequencies with minSAL, for genotypes with overall frequency >0.05, showing frequency (f) of each genotype over all populations, the correlation coefficient (R 2), gradient, and p‐value.
The MGE7243/243 homozygous genotype was also negatively correlated with minSAL (f = 0.464, R 2 = .644, p < .0001), while there was no correlation between the heterozygous genotype, MGE7243/246 and minSAL. Both MGE7246/249 (f = 0.057, R 2 = .379, p = .007) and MGE7249/249 (f = 0.0.083, R 2 = .374, p = .007) were positively correlated with minSAL (Figure 4). Although the observed frequency of the MGE7243/249 genotype was 0.009 (n = 7 mussels observed), the expected frequency was 0.13 (n = 100.6 mussels expected). The ratio of observed to expected for this genotype (0.07) was very low, even in a background of heterozygote deficiencies. The ratios of observed to expected for the three other heterozygous genotypes with expected frequency >0.05, MGE7243/246, MGE7240/243, and MGE7246/249, were 0.37, 0.50, and 0.87, respectively.
The variables minSST and minCHL‐a were included with minSAL in the top‐ranked model for the GLZ analysis based on the genetic response index f(MGE7243/243) but neither was significantly correlated with any MGE7 allele or genotype.
4. DISCUSSION
In this study, focusing on 18 populations of M. galloprovincialis along the eastern coast of the Adriatic Sea, neutral multi‐locus microsatellite analysis revealed low levels of genetic differentiation, very large population sizes and heterozygote deficiencies (consistent with a large body of work, e.g., Daniels & Litvaitis, 2017, Hedgecock et al., 2004, Wei et al., 2013b, Zouros & Foltz, 1984). These results add fine‐scale detail to the literature and are consistent with other regional or basin‐wide studies that showed either well‐connected populations in the Adriatic Sea close to panmixia, or weak east–west and/or north–south differentiation (Giantsis et al., 2014; Paterno et al., 2019; Štambuk et al., 2013; Wenne et al., 2022). We also report a pronounced gradient in allelic and genotypic frequencies at one locus that is very strongly associated with environmental variation along the eastern Adriatic Sea coastline.
The potential role of selection in generating a gradient of genetic variation is supported by the seascape analyses that revealed a strong association between the frequency of allele MGE7243 and minSAL. In the GLZ modeling, the associations between genetic variation and minSAL increased in strength as the information included in each genetic index honed in on the occurrence of the MGE7243 allele. While models using the genetic index based on neutral loci (LinF ST5) were not statistically significant and minSAL was not included in the top‐ranked model, the measures that included the MGE7 locus (LinF ST6 and LinF STMGE7) were strongly associated with the geospatial variable Lat and, to a lesser extent, minSAL. The genetic indices based solely on the frequency of the allele MGE7243 and on the homozygous genotype MGE7243/243 were strongly associated minSAL, and the models were statistically highly significant (p < .0001). The geospatial variable Lat was not statistically significant in these multivariate models, this result being consistent with the stronger single variable correlation of MGE7243 allele frequency with minSAL (R 2 = .635, p < .001) than with Lat (R 2 = .476, p = .015). The association between the MGE7243 allele and minSAL extended to the homozygous MGE7243/243 genotype, but not to the heterozygous MGE7243/X genotype. Analysis of regional subsets of the data (e.g., northern, central, southern Adriatic) and analysis with or without individual sites (e.g., farmed and wild samples from Boka Kotorska Bay) revealed that the overall relationship is robust and not dependent on the inclusion or exclusion of specific sites (results not shown). Variations in allele frequencies that coincide with changes in an environmental variable or geographic cline may arise from chance effects and historic events (Schmidt et al., 2008), but the distinctive distribution pattern of the homozygous and heterozygous genotypes is much less likely to arise by chance alone and suggests a selective pressure. The MGE7 locus may be a neutral hitchhiker locus situated close to a gene under selection or it may itself be part of the coding region of a gene under selection (Gagnaire et al., 2015). This would not be surprising, since the locus was derived from EST sequences from GenBank (Yu & Li, 2007); it was however unexpected since this locus has previously been used successfully as a neutral marker (Westfall, 2011), including in one study focused on the Croatian coast (Štambuk et al., 2013). The trend of decreasing frequency of the MGE7243 allele with increasing minimum salinity is not balanced equally by the remaining alleles. The second most common genotype, MGE7246/246, is not correlated with minSAL, while less common genotypes involving the MGE7249 allele such as MGE7246/249 and MGE7249/249 are positively correlated (R 2 = .379 and 0.374, respectively, p = .007 for both). The MGE7243 allele appears to confer an advantage in low‐salinity environments only when alleles MGE7246 and MGE7249 are not also present as the heterozygote combination, while MGE7249 may be advantageous in high‐salinity environments either as a homozygote or heterozygote. This is supported by the low frequency of the MGE7243/249 genotype which was lower than expected even in a background of heterozygote deficiency.
For the locus MGE7, we see a complex picture of relationships between the different genotypes and the environmental variable minSAL and interactions between alleles which suggest that the locus, or a nearby gene, may be under selection with two– three different alleles. A BLASTn search of the MGE7 locus against M. galloprovincialis whole genome shotgun contigs returned an alignment with contig mg10_S01094 (accession UYJE01010330.1) and a BLASTp search found 7 predicted proteins VDI82194.1 ‐ VDI82200.1. The 7 coding sequences span nucleotides 66,989–122,497 of UYJE01010330.1 with VDI82194.1 and VDI82195.1 covering the entire range. A conserved domain search retrieved 4 RNA recognition motif (RRM) superfamily domains (cl17169) close to the N‐terminal and a SPOC SHARP domain (cd21543) close to the C‐terminal. This is similar to the structure of the SHARP protein (also known as Msx2‐interacting protein) which has 3 N‐terminal RRM domains and a C‐terminal SPOC SHARP domain (Arieti et al., 2014). The BLASTp search also returned significant alignments to Msx2‐interacting proteins in other mollusks, including the mussel Mytilus coruscus, the oyster Crassostrea virginica and the abalone Haliotis rubra. Msx2 is a transcriptional regulator with a well‐documented role in the regulation of cadherins, molecules involved in cell–cell adhesion (e.g. Liang et al., 2016). Cadherins, in turn, are involved in the response to hyposalinity in the eastern oyster, Crassostrea virginica (Jones et al., 2019). While the link between MGE7 and cadherin regulation is somewhat tenuous, it does suggest a possible mechanism for our hypothesized link between MGE7 variants and salinity. However, the correlation of the variables minSAL and maxSST (R = −0.911) and consequent removal of maxSST from the seascape analyses means that firm conclusions about which of the variables is most important to the mechanism of the putative selection cannot be made. The correlation can be characterized by a progression from island sites close to the open sea (OSOR, ML, MLC) with high minimum salinity and mild summer temperatures, through to sites in enclosed bays and lagoons (CF, CW, IF, IW in Boka Kotorska Bay and SBF in Butrint Lagoon) with high freshwater inputs and/or low water exchange with the sea (low minimum salinity) and high summer temperatures (Bellafiore et al., 2011; Moisiu et al., 2016). Intermediate sites include those with low (LKF, CK, KLW, Z, U, and STF) and moderate (STW, NF, BF, and BW) levels of freshwater influence and moderate summer temperatures.
Both maxSST and minSAL are relevant to many biological processes that may affect individual survival, and have long been known to play roles in genetic selection related to environmental clines or habitat mosaics (e.g., Gardner & Palmer, 1998; Halpin et al., 2002; Hilbish & Koehn, 1985; Logan et al., 2012; Negri et al., 2013; Wenne et al., 2020). While some differences in tolerance to thermal and hypoosmotic stresses are linked to differences at the individual gene level (such as in the Lap gene of M. edulis in Long Island Sound, New York State, Hilbish et al., 1982), other differences are at the regulatory level of transcription or post‐translational modification. In an analysis of the transcriptional response of Mytilus congeners to acute heat and low‐salinity stress, of 45 genes that responded to both stressors, the response was in opposite directions for 44 of the genes, with the one gene with a response in the same direction encoding a thioredoxin reductase enzyme involved in oxidative stress (Lockwood et al., 2015). Although relatively few changes in transcriptional regulation between M. galloprovincialis and M. trossulus resulted in substantial differences in thermal tolerances between the congeners (Lockwood et al., 2015), the severe impact of heat stress on highly evolved and conserved systems such as energy metabolism and detoxification argue that there is little room for further evolutionary adaptation to thermal stress (Michaelidis et al., 2014; Tomanek, 2012).
Environmental stresses, including thermal and hyposaline stress, lead to increased levels of oxidative stress, thus if selection is the driving force behind the variation in MGE7 allele frequencies, the mechanism may be related to pathways that deal with oxidative stress rather than mechanisms specifically related to low salinity or high temperature (Hamer et al., 2008; Michaelidis et al., 2014). The effects of environmental stresses operate synergistically to increase overall levels of oxidative stress, thereby reducing tolerance to individual stresses. For example, heavy metal pollution increases sensitivity to thermal stress (Michaelidis et al., 2014). Environmental stresses increase energy demands generally in order to maintain cellular homeostasis, repair damaged proteins, and detoxify reactive oxygen species (ROS). There is a trade‐off between energy metabolism which generates ROS and expending energy to detoxify ROS, which can reach a tipping point particularly during times of thermal stress when food availability may be at its lowest (Michaelidis et al., 2014). Environmental stresses also increase susceptibility to disease which may trigger mass mortalities (Di Camillo & Cerrano, 2015; Lejeusne et al., 2010). Any adaptive mechanism that enhances the ability of mussels to respond to oxidative stress is expected to be favored in environments which experience extremes in both sea surface temperature and salinity (e.g., Boka Kotorska Bay and Butrint Lagoon, the locations of sampling sites CF, CW, IF, IW, and SBF) (Hamer et al., 2008, Michaelidis et al., 2014).
A limitation of this study is the small number of microsatellite loci available for use. Surprisingly for such an ecologically and economically important species group, there has been only a limited number of microsatellite loci developed for use with the Mytilus species complex (Araneda et al., 2016; Diz & Presa, 2008, 2009; Giantsis et al., 2014; Larraín et al., 2015), and a limited number of studies investigating population genetic structure of M. galloprovincialis using microsatellites. Interestingly, markers have often not performed well when used on mussels from different geographic regions, with, for example, only a subset of markers developed for western Mediterranean mussels (Diz & Presa, 2008), performing satisfactorily for eastern Mediterranean mussels (Giantsis et al., 2014; Štambuk et al., 2013). We suspect that this may be because of the regional differences that appear to exist in the pan‐genome of M. galloprovincialis (Gerdol et al., 2020), which also may explain the very high proportion of null alleles (i.e., via hemizygosity) that is so often reported for this species and why so many loci are out of HWE. More recently population genetic analyses have been performed using SNPs (Gardner et al., 2016; Zbawicka et al., 2014) including the Mediterranean and Black Seas (Paterno et al., 2019; Wenne et al., 2022). Our data are consistent with the SNP results but add to our understanding by being much more focussed geographically and using, where possible, local environmental data. Although SNPs are now the marker of choice (Fischer et al., 2017), microsatellites offer complementary information to SNPs, having a higher mutation rate and more recent temporal scale of inference (Fotsing et al., 2019; Waits & Storfer, 2015; Wu et al., 2021).
The findings of this study have important implications for aquaculture and conservation management in relation to climate and other anthropogenic change. Aquacultural practice along the eastern Adriatic coast is largely traditional with spat often transferred from natural spawning grounds to aquaculture sites with favorable conditions for growth (Giantsis et al., 2014; Kovačić et al., 2017). The traditional methods of mussel farming employed are extremely labor intensive so the matching of environmental conditions between the natal site of the spat and the aquaculture site is recommended to reduce selective mortality related to environmental stresses (Mandić et al., 2017; Ramón et al., 2007). This is illustrated by the MGE7 locus, with the frequency of genotypes differing between environments. Aquaculture sites are often in bays and lagoons (for example, sampling sites CF and IF in Boka Kotorska Bay and SBF in Butrint Lagoon) with high maxSST and low minSAL. Our results suggest that juveniles collected from areas of low environmental variability will have lower proportions of the favorable MGE7243 allele than juveniles collected from areas of high environmental variability and are more likely to undergo selective mortality.
Climate change has altered the geographic range of many species, moving ranges poleward and has been implicated in the success of M. galloprovincialis as an invasive species outside of its native range (reviewed by Gardner et al., 2021). Within its native range of the Mediterranean Sea, reports of mass mortalities linked to extended periods of elevated temperatures are increasing (Galli et al., 2017; Lejeusne et al., 2010; Michaelidis et al., 2014; Verdura et al., 2019). For example, widespread disease triggered by long‐lasting high water temperatures and eutrophic waters in the northern Adriatic Sea led to mortality across multiple taxa, including M. galloprovincialis, which, in some areas, decreased in abundance from ~20% coverage to ~2% in 2013, with recovery to 9.5% by 2014 (Di Camillo & Cerrano, 2015). Loss of large proportions of filter feeding animals will impact ecosystem functioning and water quality and increase the risk of eutrophication and further mass mortality events. Conservation and aquaculture practices should therefore be directed at maintaining genetic diversity and monitoring the quality of the environment to reduce the synergistic effects of pollution, eutrophication, and thermal stress (Harley et al., 2006; Lejeusne et al., 2010).
AUTHOR CONTRIBUTIONS
Joanna Hamilton: Conceptualization (equal); data curation (lead); formal analysis (equal); funding acquisition (equal); investigation (equal); methodology (equal); resources (supporting); software (equal); validation (equal); visualization (equal); writing – original draft (equal); writing – review and editing (lead). Marina Piria: Conceptualization (equal); funding acquisition (equal); methodology (equal); project administration (supporting); resources (lead); writing – review and editing (supporting). Ana Gavrilović: Resources (supporting); writing – review and editing (supporting). Mirna Mrkonjić Fuka: Resources (supporting); writing – review and editing (supporting). Lidija Svečnjak: Resources (supporting); writing – review and editing (supporting). Slađana Nikolić: Resources (supporting); writing – review and editing (supporting). Rigers Bakiu: Resources (supporting); writing – review and editing (supporting). Gardner Jonathan: Conceptualization (equal); formal analysis (equal); funding acquisition (equal); investigation (equal); methodology (equal); project administration (lead); resources (lead); software (equal); supervision (lead); validation (equal); visualization (equal); writing – original draft (equal); writing – review and editing (lead).
Supporting information
Data S1. Supporting Information.
ACKNOWLEDGMENTS
We thank Mario Lovrinov, Dragana Milošević, and Tea Tomljanović for help locating and collecting mussels, and Catarina Silva and XiangZhao Guo for help in the laboratory and with preliminary data acquisition. This work was supported by funding from the University of Zagreb Faculty of Agriculture, Croatia to MP, MF, LS, and by funding from the School of Biological Sciences, Victoria University of Wellington, New Zealand to JPAG. JSH was supported by a scholarship from VUW, NZ.
Hamilton, J. S. , Piria, M. , Gavrilović, A. , Mrkonjić Fuka, M. , Svečnjak, L. , Nikolić, S. , Bakiu, R. , & Gardner, J. P. A. (2023). Limited population genetic variation but pronounced seascape genetic structuring in populations of the Mediterranean mussel (Mytilus galloprovincialis) from the eastern Adriatic Sea. Ecology and Evolution, 00, e9729. 10.1002/ece3.9729
DATA AVAILABILITY STATEMENT
The data that support the findings of this study are openly available in Dryad at [https://doi.org/10.5061/dryad.p8cz8w9t8].
REFERENCES
- Anderson, M. J. , Gorley, R. N. , & Clarke, K. R. (2008). Permanova+ for PRIMER: Guide to software and statistical methods. PRIMER‐e. [Google Scholar]
- Araneda, C. , Noriega, X. , Jilberto, F. , Ruíz, J. , Jara, C. , Lafarga, F. , & Larraín, M. A. (2016). Development of genome‐wide microsatellite resources in a commercially important mussel species (Mytilus chilensis). Animal Genetics, 47(3), 387–388. 10.1111/age.12398 [DOI] [PubMed] [Google Scholar]
- Arieti, F. , Gabus, C. , Tambalo, M. , Huet, T. , Round, A. , & Thore, S. (2014). The crystal structure of the Split end protein SHARP adds a new layer of complexity to proteins containing RNA recognition motifs. Nucleic Acids Research, 42(10), 6742–6752. 10.1093/nar/gku277 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arribas, L. P. , Donnarumma, L. , Paloma, M. G. , & Scrosati, R. A. (2014). Intertidal mussels as ecosystem engineers: Their associated invertebrate biodiversity under contrasting wave exposures. Marine Biodiversity, 44, 203–211. 10.1007/s12526-014-0201-z [DOI] [Google Scholar]
- Artegiani, A. , Bregant, D. , Paschini, E. , Pinardi, N. , Raicich, F. , & Russo, A. (1997). The Adriatic Sea general circulation. Part I: Air‐Sea interactions and water mass structure. Journal of Physical Oceanography, 27, 1492–1514. [DOI] [Google Scholar]
- Bellafiore, D. , Guarnieri, A. , Grilli, F. , Penna, P. , Bortoluzzi, G. , Giglio, F. , & Pinardi, N. (2011). Study of the hydrodynamical processes in the Boka Kotorska Bay with a finite element model. Dynamics of Atmospheres and Oceans, 52, 298–321. 10.1016/j.dynatmoce.2011.03.005 [DOI] [Google Scholar]
- Böhm, E. , Banzon, V. , D'Acunzo, E. , D'Ortenzo, F. , & Santoleri, R. (2003). Adriatic Sea surface temperature and ocean colour variability during the MFSPP. Annales Geophysicae, 21, 137–149. 10.5194/angeo-21-137-2003 [DOI] [Google Scholar]
- Bonacci, O. , Željković, I. , & Galić, A. (2013). Karst rivers' particularity: An example from Dinaric karst (Croatia/Bosnia and Herzegovina). Environmental Earth Sciences, 70(2), 963–974. 10.1007/s12665-012-2187-9 [DOI] [Google Scholar]
- Braby, C. E. , & Somero, G. N. (2006). Following the heart: Temperature and salinity effects on heart rate in native and invasive species of blue mussels (genus Mytilus). Journal of Experimental Biology, 209, 2554–2566. 10.1242/jeb.02259 [DOI] [PubMed] [Google Scholar]
- Branch, G. M. , & Steffani, C. N. (2004). Can we predict the effects of alien species? A case‐history of the invasion of South Africa by Mytilus galloprovincialis (Lamarck). Journal of Experimental Marine Biology and Ecology, 300, 189–215. 10.1016/j.jembe.2003.12.007 [DOI] [Google Scholar]
- Bray, L. , Kassis, D. , & Hall‐Spencer, J. M. (2017). Assessing larval connectivity for marine spatial planning in the Atlantic. Marine Environmental Research, 125, 73–81. 10.1016/j.marenvres.2017.01.006 [DOI] [PubMed] [Google Scholar]
- Cárcamo, D. (2021). Does marine geospatial/environmental variation affect genetic variation? A meta‐analysis . MSc Thesis. Victoria University of Wellington. https://openaccess.wgtn.ac.nz/articles/thesis/Does_marine_geospatial_environmental_variation_affect_genetic_variation_A_meta‐analysis/16573280/1
- Da Ros, L. , Bressan, M. , & Marin, M. G. (1985). Reproductive cycle of the mussel (Mytilus galloprovincialis Lmk) in Venice lagoon (North Adriatic). Italian Journal of Zoology, 52(3–4), 223–229. 10.1080/11250008509440522 [DOI] [Google Scholar]
- Daniels, E. Q. , & Litvaitis, M. K. (2017). Does peripheral dislodgement contribute to heterozygote deficiencies in blue mussels? Invertebrate Biology, 136(3), 301–308. 10.1111/ivb.12183 [DOI] [Google Scholar]
- Di Camillo, C. G. , & Cerrano, C. (2015). Mass mortality events in the NW Adriatic Sea: Phase shift from slow‐ to fast‐ growing organisms. PLoS One, 10(5), e0126689. 10.1371/journal.pone.0126689 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diz, A. P. , & Presa, P. (2008). Regional patterns of microsatellite variation in Mytilus galloprovincialis from the Iberian Peninsula. Marine Biology, 154, 227–286. 10.1007/s00227-008-0921-3 [DOI] [Google Scholar]
- Diz, A. P. , & Presa, P. (2009). The genetic diversity pattern of Mytilus galloprovincialis in Galician Rías (NW Iberian estuaries). Aquaculture, 287, 278–285. 10.1016/j.aquaculture.2008.10.029 [DOI] [Google Scholar]
- Do, C. , Waples, R. S. , Peel, D. , Macbeth, G. M. , Tillett, B. J. , & Ovenden, J. R. (2014). NeEstimator V2: re‐implementation of software for the estimation of contemporary effective population size (N e ) from genetic data. Molecular Ecology Resources, 14(1), 209–214. 10.1111/1755-0998.12157 [DOI] [PubMed] [Google Scholar]
- Dubois, M. , Rossi, V. , Ser‐Giacomi, E. , Arnaud‐Haond, S. , López, C. , & Hernández‐García, E. (2016). Linking basin‐scale connectivity, oceanography and population dynamics for the conservation and management of marine ecosystems. Global Ecology and Biogeography, 25, 503–515. 10.1111/geb.12431 [DOI] [Google Scholar]
- Đurović, B. , Đurović, I. , Joksimović, A. , Crnojević, V. , Đukanović, S. , & Pestorić, B. (2018). Monitoring the eutrophication using Landsat 8 in the Boka Kotorska Bay. Acta Adriatica, 59(1), 17–34. 10.32582/aa.59.1.2 [DOI] [Google Scholar]
- Earl, D. A. , & von Holdt, B. M. (2012). Structure harvester: A website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genetic Resources, 4, 359–361. 10.1007/s12686-011-9548-7 [DOI] [Google Scholar]
- Estournel, C. , Marsaleix, P. , & Ulses, C. (2021). A new assessment of the circulation of Atlantic and intermediate waters in the eastern Mediterranean. Progress in Oceanography, 198, 102673. 10.1016/j.pocean.2021.102673 [DOI] [Google Scholar]
- Evanno, G. , Regnaut, S. , & Goudet, J. (2005). Detecting the number of clusters of individuals using the software STRUCTURE: A simulation study. Molecular Ecology, 14, 2611–2620. 10.1111/j.1365-294X.2005.02553.x [DOI] [PubMed] [Google Scholar]
- Evans, T. G. , & Somero, G. N. (2010). Phosphorylation events catalyzed by major cell Signalling proteins differ in response to thermal and osmotic stress among native (Mytilus californianus and Mytilus trossulus) and invasive (Mytilus galloprovincialis) species of mussels. Physiological and Biochemical Zoology, 83(6), 984–996. 10.1086/656192 [DOI] [PubMed] [Google Scholar]
- Excoffier, L. , & Lischer, H. E. (2010). Arlequin suite ver 3.5: A new series of programs to perform population genetic analyses under Linux and windows. Molecular Ecology Resources, 10, 564–567. 10.1111/j.1755-0998.2010.02847.x [DOI] [PubMed] [Google Scholar]
- Excoffier, L. , Smouse, P. E. , & Quattro, J. M. (1992). Analysis of molecular variance inferred from metric distances among DNA haplotypes: Application to human mitochondrial DNA restriction data. Genetics, 131, 479–491. 10.1093/genetics/131.2.479 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fischer, M. C. , Rellstab, C. , Leuzinger, M. , Roumet, M. , Gugerli, F. , Shimizu, K. K. , Holderegger, R. , & Widmer, A. (2017). Estimating genomic diversity and population differentiation – An empirical comparison of microsatellite and SNP variation in Arabidopsis halleri . BMC Genomics, 18, 69. 10.1186/s12864-016-3459-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fotsing, S. F. , Margoliash, J. , Wang, C. , Saini, S. , Yanicky, R. , Shleizer‐Burko, S. , Goren, A. , & Gymrek, M. (2019). The impact of short tandem repeat variation on gene expression. Nature Genetics, 51, 1652–1659. 10.1038/s41588-019-0521-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gaggiotti, O. E. , Bekkevold, D. , Jørgensen, H. B. H. , Foll, M. , Carvalho, G. R. , Andre, C. , & Ruzzante, D. E. (2009). Disentangling the effects of evolutionary, demographic, and environmental factors influencing genetic structure of natural populations: Atlantic herring as a case study. Evolution, 63(11), 2939–2951. 10.1111/j.1558-5646.2009.00779.x [DOI] [PubMed] [Google Scholar]
- Gagnaire, P.‐A. , Broquet, T. , Aurelle, D. , Viard, F. , Souissi, A. , Bonhomme, F. , Arnaud‐Haond, S. , & Bierne, N. (2015). Using neutral, selected, and hitchhiker loci to assess connectivity of marine populations in the genomic era. Evolutionary Applications, 8, 769–786. 10.1111/eva.12288 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Galli, G. , Solidoro, C. , & Lovato, T. (2017). Marine heat waves Hazard 3D maps and the risk for low motility organisms in a warming Mediterranean Sea. Frontiers in Marine Science, 4, 136. 10.3389/fmars.2017.00136 [DOI] [Google Scholar]
- Gao, X. , & Starmer, J. D. (2008). AWclust: Point‐and‐click software for non‐parametric population structure analysis. BMC Bioinformatics, 9, 77. 10.1186/1471-2105-9-77 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gardeström, J. , Pereyra, R. T. , & André, C. (2008). Characterization of six microsatellite loci in the Baltic blue mussel Mytilus trossulus and cross‐species amplification in North Sea Mytilus edulis . Conservation Genetics, 9, 1003–1005. 10.1007/s10592-007-9432-x [DOI] [Google Scholar]
- Gardner, J. P. A. , Oyarzún, P. A. , Toro, J. E. , Wenne, R. , & Zbawicka, M. (2021). Phylogeography of southern hemisphere blue mussels of the genus Mytilus: Evolution, biosecurity, aquaculture and food labelling. Oceanography and Marine Biology: Annual Reviews, 59, 139–232. 10.1201/9781003138846-3 [DOI] [Google Scholar]
- Gardner, J. P. A. , & Palmer, N. L. (1998). Size‐dependent, spatial and temporal genetic variation at a leucine aminopeptidase (LAP) locus among blue mussel (Mytilus galloprovincialis) populations along a salinity gradient. Marine Biology, 132, 275–281. 10.1007/s002270050393 [DOI] [Google Scholar]
- Gardner, J. P. A. , Zbawicka, M. , Westfall, K. M. , & Wenne, R. (2016). Invasive blue mussels threaten regional scale genetic diversity in mainland and remote offshore locations: The need for baseline data and enhanced protection in the Southern Ocean. Global Change Biology, 22, 3182–3195. 10.1111/gcb.13332 [DOI] [PubMed] [Google Scholar]
- Geller, J. B. (1999). Decline of a native mussel masked by sibling species invasion. Conservation Biology, 13, 661–664. 10.1046/j.1523-1739.1999.97470.x [DOI] [Google Scholar]
- Gerdol, M. , Moreira, R. , Cruz, F. , Gómez‐Garrido, J. , Vlasova, A. , Rosani, U. , Venier, P. , Naranjo‐Ortiz, M. A. , Murgarella, M. , Greco, S. , Balseiro, P. , Corvelo, A. , Frias, L. , Gut, M. , Gabaldón, T. , Pallavicini, A. , Canchaya, C. , Novoa, B. , Alioto, T. S. , … Figueras, A. (2020). Massive gene presence‐absence variation shapes an open pan‐genome in the Mediterranean mussel. Genome Biology, 21, 275. 10.1186/s13059-020-02180-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giantsis, I. A. , Mucci, N. , Randi, E. , Abatzopoulos, T. J. , & Apostolidis, A. P. (2014). Microsatellite variation of mussels (Mytilus galloprovincialis) in central and eastern Mediterranean: Genetic panmixia in the Aegean and the Ionian seas. Journal of the Marine Biological Association of the United Kingdom, 94(4), 797–809. 10.1017/S0025315414000174 [DOI] [Google Scholar]
- Gruber, M. A. M. , Hoffmann, B. D. , Ritchie, P. A. , & Lester, P. J. (2013). The conundrum of the yellow crazy ant (Anoplolepis gracilipes) reproductive mode: no evidence for dependent lineage genetic caste determination. Insectes Sociaux, 60, 135–145. 10.1007/s00040-012-0277-z [DOI] [Google Scholar]
- Halpin, P. M. , Sorte, C. J. , Hofmann, G. E. , & Menge, B. A. (2002). Patterns of variation in levels of Hsp70 in natural rocky shore populations from microscales to mesoscales. Integrative and Comparative Biology, 42(4), 815–824. 10.1093/icb/42.4.815 [DOI] [PubMed] [Google Scholar]
- Hamer, B. , Jakšić, Ž. , Pavičić‐Hamer, D. , Perić, L. , Medaković, D. , Ivanković, D. , Pavičić, J. , Zilberberg, C. , Schröder, H. C. , Müller, W. E. G. , Smodlaka, N. , & Batel, R. (2008). Effect of hypoosmotic stress by low salinity acclimation of Mediterranean mussels Mytilus galloprovincialis on biological parameters used for pollution assessment. Aquatic Toxicology, 89, 137–151. 10.1016/j.aquatox.2008.06.015 [DOI] [PubMed] [Google Scholar]
- Hamilton, J. S. (2019). Using neutral and outlier loci to uncover the genetic structure of Mytilus galloprovincialis populations from the eastern Adriatic coast using population genetic and seascape genetic approaches. MSc thesis. Victoria University of Wellington. http://researcharchive.vuw.ac.nz/handle/10063/8161 [Google Scholar]
- Harley, C. D. G. , Hughes, A. R. , Hultgren, K. M. , Miner, B. G. , Sorte, C. J. B. , Thomber, C. S. , Rodriguez, L. F. , Tomanek, L. , & Williams, S. L. (2006). The impacts of climate change in coastal marine systems. Ecology Letters, 9, 228–241. 10.1111/j.1461-0248.2005.00871.x [DOI] [PubMed] [Google Scholar]
- Hedgecock, D. , Li, G. , Hubert, S. , Bucklin, K. , & Ribes, V. (2004). Widespread null alleles and poor cross‐species amplification of microsatellite DNA loci cloned from the Pacific oyster, Crassostrea gigas . Journal of Shellfish Research, 23(2), 379–385. [Google Scholar]
- Hilbish, T. J. , Deaton, L. E. , & Koehn, R. K. (1982). Effect of an allozyme polymorphism on regulation of cell volume. Nature, 298, 688–689. 10.1038/298688a0 [DOI] [PubMed] [Google Scholar]
- Hilbish, T. J. , & Koehn, R. K. (1985). Dominance in physiological phenotypes and fitness at an enzyme locus. Science, 229, 52–54. 10.1126/science.4012310 [DOI] [PubMed] [Google Scholar]
- Hill, W. G. (1981). Estimation of effective population size from data on linkage disequilibrium. Genetical Research Cambridge, 38, 209–216. 10.1017/S0016672300020553 [DOI] [Google Scholar]
- Holderegger, R. , Kamm, U. , & Gugerli, F. (2006). Adaptive vs. neutral genetic diversity: Implications for landscape genetics. Landscape Ecology, 21, 797–807. 10.1007/s10980-005-5245-9 [DOI] [Google Scholar]
- IOR . (2022). Database of the sea bathing water quality in Croatia. Developed by Institute for Oceanography and Fisheries. Retrieved October, 2, 2010 from: https://vrtlac.izor.hr/ords/kakvoca/kakvoca_detalji10
- Jones, H. R. , Johnson, K. M. , & Kelly, M. W. (2019). Synergistic effects of temperature and salinity on the gene expression and physiology of Crassostrea virginica . Integrative and Comparative Biology, 59(2), 306–319. 10.1093/icb/icz035 [DOI] [PubMed] [Google Scholar]
- Kalinowski, S. T. (2005). HP‐RARE 1.0: A computer program for performing rarefaction on measures of allelic richness. Molecular Ecology Notes, 5, 187–189. 10.1111/j.1471-8286.2004.00845.x [DOI] [Google Scholar]
- Kearse, M. , Moir, R. , Wilson, A. , Stones‐Havas, S. , Cheung, M. , Sturrock, S. , Buxton, S. , Cooper, A. , Markowitz, S. , Duran, C. , Thierer, T. , Ashton, B. , Meintjes, P. , & Drummond, A. (2012). Geneious basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics Methods and Protocols, 28, 1647–1649. 10.1093/bioinformatics/bts199 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovačić, I. , Pavičić‐Hamer, D. , Kanduč, T. , & Hamer, B. (2017). Adaptation of cultured mussel Mytilus galloprovincialis Lamarck, 1819 from the northern Adriatic Sea to nearby aquaculture sites and translocation. Acta Adriatica, 58(2), 285–296. 10.32582/aa.58.2.8 [DOI] [Google Scholar]
- Krivokapić, S. , Pestorić, B. , Bosak, S. , Kušpilić, G. , & Wexels Riser, C. (2011). Trophic state of Boka Kotorska Bay (South‐Eastern Adriatic Sea). Fresenius Environmental Bulletin, 20(8), 1960–1969. [Google Scholar]
- Larraín, M. A. , Díaz, N. F. , Lamas, C. , Uribe, C. , Jilberto, F. , & Araneda, C. (2015). Heterologous microsatellite‐based genetic diversity in blue mussel (Mytilus chilensis) and differentiation among localities in southern Chile. Latin American Journal of Aquatic Research, 43(5), 998–1010. 10.3856/vol43-issue5-fulltext-20 [DOI] [Google Scholar]
- Lejeusne, C. , Chevaldonné, P. , Pergent‐Martini, C. , Boudouresque, C. F. , & Pérez, T. (2010). Climate change effects on a miniature ocean: The highly diverse, highly impacted Mediterranean Sea. Trends in Ecology and Evolution, 25(4), 250–260. 10.1016/j.tree.2009.10.009 [DOI] [PubMed] [Google Scholar]
- Liang, H. , Zhang, Q. , Lu, J. , Yang, G. , Tian, N. , Wang, X. , Tan, Y. , & Tan, D. (2016). MSX2 induces trophoblast invasion in human placenta. PLoS One, 11(4), e0153656. 10.1371/journal.pone.0153656 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lipizer, M. , Partescano, E. , Rabitti, A. , Giorgetti, A. , & Crise, A. (2014). Qualified temperature, salinity and dissolved oxygen climatologies in a changing Adriatic Sea. Ocean Science, 10, 771–797. 10.5194/os-10-771-2014 [DOI] [Google Scholar]
- Lockwood, B. L. , Connor, K. M. , & Gracey, A. Y. (2015). The environmentally tuned transcriptomes of Mytilus mussels. The Journal of Experimental Biology, 218, 1822–1833. 10.1242/jeb.118190 [DOI] [PubMed] [Google Scholar]
- Logan, C. A. , Kost, L. E. , & Somero, G. N. (2012). Latitudinal differences in Mytilus californianus thermal physiology. Marine Ecology Progress Series, 450, 93–105. 10.3354/meps09491 [DOI] [Google Scholar]
- Lupo, C. , Bougeard, S. , Le Bihan, V. , Blin, J. L. , Allain, G. , Azéma, P. , Benoit, F. , Béchemin, C. , Bernard, I. , Blachier, P. , Brieau, L. , Danion, M. , Garcia, A. , Gervasoni, E. , Glize, P. , Lainé, A. , Lapègue, S. , Mablouké, C. , Poirier, L. , … Le Bouquin, S. (2021). Mortality of marine mussels Mytilus edulis and M. galloprovincialis: Systematic literature review of risk factors and recommendations for future research. Reviews in Aquaculture, 13(1), 504–536. 10.1111/raq.12484 [DOI] [Google Scholar]
- Mandić, M. , Ikica, Z. , & Gvozdenović, S. (2017). Mariculture in the Boka Kotorska Bay: Tradition, current state and perspective. In Joksimović A., Djurović M., Semenov A. V., Zonn I. S., & Kostianoy A. G. (Eds.), The Boka Kotorska Bay Environment (pp. 395–409). Springer. 10.1007/698_2016_33 [DOI] [Google Scholar]
- Manel, S. , Schwartz, M. K. , Luikart, G. , & Taberlet, P. (2003). Landscape genetics: Combining landscape ecology and population genetics. Trends in Ecology and Evolution, 18(4), 189–197. 10.1016/S0169-5347(03)00008-9 [DOI] [Google Scholar]
- Michaelidis, B. , Pörtner, H.‐O. , Sokolova, I. , & Tomanek, L. (2014). Advances in predicting the impacts of global warming on the mussels Mytilus galloprovincialis in the Mediterranean Sea. In Goffredo S. & Dubinsky Z. (Eds.), The Mediterranean Sea: Its history and present challenges (pp. 319–339). Springer. 10.1007/978-94-007-6704-1_18 [DOI] [Google Scholar]
- Moisiu, L. , Panagiotopoulos, I. P. , Durmishi, Ҫ. , Kapsimalis, V. , & Anagnostou, C. (2016). The anoxic Butrint Lagoon, SW Albania. Environmental Earth Sciences, 75, 1443. 10.1007/s12665-016-6259-0 [DOI] [Google Scholar]
- Negri, A. , Oliveri, C. , Sforzini, S. , Mignione, F. , Viarengo, A. , & Banni, M. (2013). Transcriptional response of the mussel Mytilus galloprovincialis (lam.) following exposure to heat stress and copper. PLoS One, 8(6), e66802. 10.1371/journal.pone.0066802 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oliver, E. C. J. , Donat, M. G. , Burrows, M. T. , Moore, P. J. , Smale, D. A. , Alexander, L. V. , Benthuysen, J. A. , Feng, M. , Gupta, A. S. , Hobday, A. J. , Holbrook, N. J. , Perkins‐Kirkpatrick, S. E. , Scannell, H. A. , Straub, S. C. , & Wernberg, T. (2018). Longer and more frequent marine heatwaves over the past century. Nature Communications, 9, 1324. 10.1038/s41467-018-03732-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pascual, M. , Rives, B. , Schunter, C. , & MacPherson, E. (2017). Impact of life history traits on gene flow: A multispecies systematic review across oceanographic barriers in the Mediterranean Sea. PLoS One, 12(5), e0176419. 10.1371/journal.pone.0176419 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Paterno, M. , Bat, L. , Souissi, J. B. , Boscari, E. , Chassanite, A. , Congiu, L. , Guarnieri, G. , Kruschel, C. , Mačić, V. , Marino, I. A. M. , Micu, D. , Milchakova, N. , Panayotova, M. , Papetti, C. , Planes, S. , Strungaru, S. , Todorova, V. R. , Voutsinas, E. , & Zane, L. (2019). A genome‐wide approach to the phylogeography of the mussel Mytilus galloprovincialis in the Adriatic and the black seas. Frontiers in Marine Science, 6, 566. 10.3389/fmars.2019.00566 [DOI] [Google Scholar]
- Peakall, R. , & Smouse, P. E. (2012). GenAlEx 6.5: Genetic analysis in excel. Population genetic structure for teaching and research – An update. Bioinformatics, 28, 2537–2539. 10.1093/bioinformatics/bts460 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pechenik, J. A. (1999). On the advantages and disadvantages of larval stages in benthic marine invertebrate life cycles. Marine Ecology Progress Series, 177, 269–297. 10.3354/meps177269 [DOI] [Google Scholar]
- Presa, P. , Pérez, M. , & Diz, A. P. (2002). Polymorphic microsatellite markers for blue mussels (Mytilus spp.). Conservation Genetics, 3, 441–443. 10.1023/A:1020571202907 [DOI] [Google Scholar]
- Pritchard, J. K. , Stephens, M. , & Donnelly, P. (2000). Inference of population structure using multilocus genotype data. Genetics, 155, 945–959. 10.1093/genetics/155.2.945 [DOI] [PMC free article] [PubMed] [Google Scholar]
- R Core Team . (2014). R: A language and environment for statistical computing. R Foundation for Statistical Computing; http://www.R‐project.org/ [Google Scholar]
- Ramón, M. , Fernández, M. , & Galimany, E. (2007). Development of mussel (Mytilus galloprovincialis) seed from two different origins in a semi‐enclosed Mediterranean Bay (N.E. Spain). Aquaculture, 264, 148–159. 10.1016/j.aquaculture.2006.11.014 [DOI] [Google Scholar]
- Riginos, C. , & Liggins, L. (2013). Seascape genetics: Populations, individuals, and genes marooned and adrift. Geography Compass, 7(3), 197–216. 10.1111/gec3.12032 [DOI] [Google Scholar]
- Rioux Paquette, S. (2011). PopGenKit: Useful function for (batch) file conversion and data resampling in microsatellite datasets. R package.
- Rousset, F. (2008). GENEPOP ‘007: A complete re‐implementation of the GENEPOP software for Windows and Linux. Molecular Ecology Resources, 8(1), 103–106. 10.1111/j.1471-8286.2007.01931.x [DOI] [PubMed] [Google Scholar]
- Saarman, N. P. , Kober, K. M. , Simison, W. B. , & Pogson, G. H. (2017). Sequence‐based analysis of thermal adaptation and protein energy landscapes in an invasive blue mussel (Mytilus galloprovincialis). Genome Biology and Evolution, 9(10), 2739–2751. 10.1093/gbe/evx190 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schmidt, P. S. , Serrão, E. A. , Pearson, G. A. , Riginos, C. , Rawson, P. D. , Hilbish, T. J. , Brawley, S. H. , Trussell, G. C. , Carrington, E. , Wethey, D. S. , Grahame, J. W. , Bonhomme, F. , & Rand, D. M. (2008). Ecological genetics In the North Atlantic: Environmental gradients and adaptation At specific loci. Ecology, 89(11), S91–S107. 10.1890/07-1162.1 [DOI] [PubMed] [Google Scholar]
- Schneider, K. R. , & Helmuth, B. (2007). Spatial variability in habitat temperature may drive patterns of selection between an invasive and native mussel species. Marine Ecology Progress Series, 339, 157–167. 10.3354/meps339157 [DOI] [Google Scholar]
- SeaDataNet.org . Originally accessed at http://sdn.oceanbrowser.net/web‐vis/ and http://gher‐diva.phys.ulg.ac.be/emodnet/ accessed 28/10/2018. Now available at https://www.seadatanet.org/Products#/metadata/2a2aa0c5‐4054‐4a62‐a18b‐3835b304fe64.
- Segovia, N. I. , González‐Wevar, C. A. , & Haye, P. A. (2020). Signatures of local adaptation in the spatial genetic structure of the ascidian Pyura chilensis along the Southeast Pacific coast. Scientific Reports, 10, 14098. 10.1038/s41598-020-70798-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Selkoe, K. A. , D'Aloia, C. C. , Crandall, E. D. , Iacchei, M. , Liggins, L. , Puritz, J. B. , von der Heyden, S. , & Toonen, R. J. (2016). A decade of seascape genetics: Contributions to basic and applied marine connectivity. Marine Ecology Progress Series, 554, 1–19. 10.3354/meps11792 [DOI] [Google Scholar]
- Selkoe, K. A. , Gaines, S. D. , Caselle, J. E. , & Warner, R. R. (2006). Current shifts and kin aggregation explain genetic patchiness in fish recruits. Ecology, 87, 3082–3094. 10.1890/0012-9658(2006)87[3082:csakae]2.0.co;2 [DOI] [PubMed] [Google Scholar]
- Selkoe, K. A. , Scribner, K. T. , & Galindo, H. M. (2015). Waterscape genetics: Applications of landscape genetics to rivers, lakes, and seas. In Balkenhol N., Cushman S. A., Storfer A. T., & Waits L. P. (Eds.), Landscape genetics (pp. 220–246). Wiley Blackwell, West Sussex. 10.1002/9781118525258.ch13 [DOI] [Google Scholar]
- Silva, C. N. S. , & Gardner, J. P. A. (2016). Identifying environmental factors associated with the genetic structure of the New Zealand scallop: Linking seascape genetics and ecophysiological tolerance. ICES Journal of Marine Science, 73(7), 1925–1934. 10.1093/icesjms/fsv240 [DOI] [Google Scholar]
- Štambuk, A. , Šrut, M. , Šatović, Z. , Tkalec, M. , & Klobučar, G. I. V. (2013). Gene flow vs. pollution pressure: Genetic diversity of Mytilus galloprovincialis in eastern Adriatic. Aquatic Toxicology, 136‐137, 22–31. 10.1016/j.aquatox.2013.03.017 [DOI] [PubMed] [Google Scholar]
- Tomanek, L. (2012). Environmental proteomics of the mussel Mytilus: Implications for tolerance to stress and change in limits of biogeographical ranges in response to climate change. Integrative and Comparative Biology, 52(5), 648–664. 10.1093/icb/ics114 [DOI] [PubMed] [Google Scholar]
- Tomanek, L. , & Helmuth, B. (2002). Physiological ecology of rocky intertidal organisms: A synergy of concepts. Integrative and Comparative Biology, 42, 771–775. 10.1093/icb/42.4.771 [DOI] [PubMed] [Google Scholar]
- UNEP/MAP‐RAC/SPA . (2015). Adriatic Sea: Description of the ecology and identification of the areas that may deserve to be protected. By Cerrano C. Edited by Cebrian D and Requena S. RAC/SPA, Tunis. [Google Scholar]
- Van Oosterhout, C. , Hutchinson, W. F. , Wills, D. P. M. , & Shipley, P. (2004). Micro‐checker: Software for identifying and correcting genotyping errors in microsatellite data. Molecular Ecology Notes, 4, 535–538. 10.1111/j.1471-8286.2004.00684.x [DOI] [Google Scholar]
- Vaux, F. , Bohn, S. , Hyde, J. R. , & O'Malley, K. G. (2021). Adaptive markers distinguish north and South Pacific albacore amid low population differentiation. Evolutionary Applications, 14(5), 1343–1364. 10.1111/eva.13202 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Verdura, J. , Linares, C. , Ballesteros, E. , Coma, R. , Uriz, M. J. , Bensoussan, N. , & Cebrian, E. (2019). Biodiversity loss in a Mediterranean ecosystem due to an extreme warming event unveils the role of an engineering gorgonian species. Scientific Reports, 9(1), 5911. 10.1038/s41598-019-41929-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Verhoeven, K. J. F. , Simonsen, K. L. , & McIntyre, L. M. (2005). Implementing false discovery rate control: Increasing your power. Oikos, 3, 643–647. 10.1111/j.0030-1299.2005.13727.x [DOI] [Google Scholar]
- Waits, L. P. , & Storfer, A. (2015). Basics of population genetics: Quantifying neutral and adaptive genetic variation for landscape genetic studies. In Balkenhol N., Cushman S. A., Storfer A. T., & Waits L. P. (Eds.), Landscape genetics (pp. 35–57). Wiley Blackwell, West Sussex. 10.1002/9781118525258.ch03 [DOI] [Google Scholar]
- Waples, R. S. (2006). A bias correction for estimates of effective population size based on linkage disequilibrium at unlinked loci. Conservation Genetics, 7(2), 167–184. 10.1007/s10592-005-9100-y [DOI] [Google Scholar]
- Waples, R. S. , & Do, C. (2010). Linkage disequilibrium estimates of contemporary N e using highly variable genetic markers: A largely untapped resource for applied conservation and evolution. Evolutionary Applications, 3(3), 244–262. 10.1111/j.1752-4571.2009.00104.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wei, K.‐J. , Wood, A. R. , & Gardner, J. P. A. (2013a). Seascape genetics of the New Zealand greenshell mussel: Sea surface temperature explains macrogeographic scale genetic variation. Marine Ecology Progress Series, 477, 107–121. 10.3354/meps10158 [DOI] [Google Scholar]
- Wei, K.‐J. , Wood, A. R. , & Gardner, J. P. A. (2013b). Population genetic variation in the New Zealand greenshell mussel: Locus‐dependent conflicting signals of weak structure and high gene flow balanced against pronounced structure and high self‐recruitment. Marine Biology, 160, 931–949. 10.1007/s00227-012-2145-9 [DOI] [Google Scholar]
- Wenne, R. , Zbawicka, M. , Bach, L. , Strelkov, P. , Gantsevich, M. , Kuklinski, P. , Kijewski, T. , McDonald, J. H. , Sundsaasen, K. K. , Árnyasi, M. , Lien, S. , Kaasik, A. , Herkül, K. , & Kotta, J. (2020). Trans‐Atlantic distribution and introgression as inferred from single nucleotide polymorphism: Mussels mytilus and environmental factors. Genes, 11(5), 530. 10.3390/genes11050530 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wenne, R. , Zbawicka, M. , Prądzińska, A. , Kotta, J. , Herkül, K. , Gardner, J. P. A. , Apostolidis, A. P. , Poćwierz‐Kotus, A. , Rouane‐Hacene, O. , Korrida, A. , Dondero, F. , Baptista, M. , Reizopoulou, S. , Hamer, B. , Sundsaasen, K. K. , Árnyasi, M. , & Kent, M. P. (2022). Molecular genetic differentiation of native populations of Mediterranean blue mussels, Mytilus galloprovincialis (Lamarck, 1819), and the relationship with environmental variables. European Zoological Journal, 89, 755–784. [Google Scholar]
- Westfall, K. M. (2011). Molecular ecology and systematics of blue mussels (genus Mytilus) (Mytilidae; Bivalvia; Mollusca) in the southern hemisphere. PhD thesis. Victoria University of Wellington. [Google Scholar]
- White, T. A. , Stamford, J. , & Hoelzel, A. R. (2010). Local selection and population structure in a deep‐sea fish, the roundnose grenadier (Coryphaenoides rupestris). Molecular Ecology, 19, 216–226. 10.1111/j.1365-294X.2009.04446.x [DOI] [PubMed] [Google Scholar]
- Wu, Z. , Gong, H. , Zhang, M. , Tong, X. , Ai, H. , Xiao, S. , Perez‐Enciso, M. , Yang, B. , & Huang, L. (2021). A worldwide map of swine short tandem repeats and their associations with evolutionary and environmental adaptations. Genetics Selection Evolution, 53(1), 39. 10.1186/s12711-021-00631-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu, H. , & Li, Q. (2007). Development of EST‐SSRs in the Mediterranean mussel, Mytilus galloprovincialis . Molecular Ecology Notes, 7, 1308–1310. 10.1111/j.1471-8286.2007.01865.x [DOI] [Google Scholar]
- Zbawicka, M. , Wenne, R. , & Burzyński, A. (2014). Mitogenomics of recombinant mitochondrial genomes of Baltic Sea Mytilus mussels. Molecular Genetics and Genomics, 289, 1275–1287. 10.1007/s00438-014-0888-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zeng, C. , Rowden, A. A. , Clark, M. R. , & Gardner, J. P. A. (2020). Species‐specific genetic variation in response to deep‐sea environmental variation amongst vulnerable marine ecosystem indicator taxa. Scientific Reports, 10, 2844. 10.1038/s41598-020-59210-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zonn, I. S. , & Kostianoy, A. G. (2017). The Adriatic Sea. In Joksimović A., Djurović M., Semenov A. V., Zonn I. S., & Kostianoy A. G. (Eds.), The Boka Kotorska Bay Environment (pp. 19–41). Springer. 10.1007/698_2016_42 [DOI] [Google Scholar]
- Zouros, E. , & Foltz, D. W. (1984). Possible explanations of heterozygote deficiency in bivalve Molluscs. Malacologia, 25(2), 583–591. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data S1. Supporting Information.
Data Availability Statement
The data that support the findings of this study are openly available in Dryad at [https://doi.org/10.5061/dryad.p8cz8w9t8].
