Survey
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Coral Reefs (2012) 31:1135–1148 DOI 10.1007/s00338-012-0937-5 REPORT Fine-scale spatial genetic structure and clonal distribution of the cold-water coral Lophelia pertusa M. P. Dahl • R. T. Pereyra • T. Lundälv C. André • Received: 3 November 2011 / Accepted: 16 July 2012 / Published online: 2 August 2012 Ó Springer-Verlag 2012 Abstract Determining the spatial genetic structure within and among cold-water coral populations is crucial to understanding population dynamics, assessing the resilience of cold-water coral communities and estimating genetic effects of habitat fragmentation for conservation. The spatial distribution of genetic diversity in natural populations depends on the species’ mode of reproduction, and coral species often have a mixed strategy of sexual and asexual reproduction. We describe the clonal architecture of a cold-water coral reef and the fine-scale population genetic structure (\35 km) of five reef localities in the NE Skagerrak. This study represents the first of this type of analysis from deep waters. We used thirteen microsatellite loci to estimate gene flow and genotypic diversity and to describe the fine-scale spatial distribution of clonal individuals of Lophelia pertusa. Within-population genetic diversity was high in four of the five reef localities. These four reefs constitute a genetic cluster with asymmetric gene flow that indicates metapopulation dynamics. One locality, the Säcken reef, was genetically isolated and depauperate. Asexual reproduction was found to be a highly important mode of reproduction for L. pertusa: 35 genetic individuals Communicated by Biology Editor Dr. Ruth Gates Electronic supplementary material The online version of this article (doi:10.1007/s00338-012-0937-5) contains supplementary material, which is available to authorized users. M. P. Dahl (&) R. T. Pereyra C. André Department of Marine Ecology-Tjärnö, University of Gothenburg, 452 96 Strömstad, Sweden e-mail: [email protected] T. Lundälv Sven Lovén Centre of Marine Sciences-Tjärnö, University of Gothenburg, Strömstad, Sweden were found on the largest reef, with the largest clone covering an area of nearly 300 m2. Keywords Cold-water coral Clonality Spatial genetic structure Genotypic diversity Connectivity Conservation genetics Introduction Genetic diversity within and among populations is influenced by the species’ mode of reproduction. Sexual reproduction resulting in recombination increases genetic diversity within populations, while dispersal of larvae connects populations. In contrast, clonal reproduction (asexual), which lacks sexual recombination, may decrease diversity, potentially hampering adaptation to environmental change (Lasker and Coffroth 1999). On the other hand, clonal propagation allows organisms to produce progeny without sexual reproduction and thus enables species to persist when unable to complete the sexual reproductive life cycle (Honnay and Bossuyt 2005). Clonal propagation also allows genetic individuals to spread out by clonal growth and to monopolize resources locally (Pan and Price 2002). Nevertheless, all species that are considered clonal have some level of genetic recombination (Bengtsson 2003; de Meeûs et al. 2007; Gladychev et al. 2008). The degree of clonality has been shown to influence the genetic structure within and among populations by reducing effective population size and altering gene flow (Loveless and Hamrick 1984; Baums et al. 2006; Whitaker 2006). However, the importance of clonal reproduction for the spatial distribution of genetic diversity in natural populations of cold-water corals is largely unknown. Clonal establishment benefits species with a mixed strategy of sexual and asexual reproduction particularly in 123 1136 unfavourable local environments where occasional sexual recombination keeps genetic diversity higher and decreases likelihood of extinction (Bengtsson 2003). In some species, sexual reproduction and asexual reproduction contribute equally to population growth, while in others, one reproductive mode dominates. The relative contribution of sexual and asexual reproduction can vary among populations of a single species in response to biotic and abiotic factors, for example latitude (Hoffman 1986; Dorken and Eckert 2001) and population density (van Kleunen et al. 2001). The extent of clonality tends to increase at the limits of geographical distribution of the species compared to those populations located in the centre of the range and also tends to increase with population age (Lesica and Allendorf 1995; Eckert 2002; Silvertown 2008). Population connectivity occurs through the larval stage, which is the only long-distance dispersal phase. Marine larvae are, however, notoriously difficult to track in the natural environment, and genetic markers are therefore often used to infer population connectivity. In general, marine populations are believed to be highly connected, resulting in weak genetic structures. But this view has been recently challenged (Cowen et al. 2000; Hauser and Carvalho 2008). Indeed, corals show a wide range of genetic structures, from panmixia over 7,000 km (Takabayashi et al. 2003) to locally subdivided populations (see van Oppen and Gates (2006) for a review). Within coral reefs, genotypic richness depends on the frequency of larval replenishment, recruitment dynamics and longevity of genets (Eriksson 1993). The term ‘ramet’ is used for a physiologically distinct colony, and the term ‘genet’ or clone describes all ramets that are genetically identical (sensu Harper 1977). Genotypic richness is expected to decrease over time due to elimination of genets by intraspecific competition, selection and stochastic effects (Hartnett and Bazzaz 1985; Eriksson 1989). However, genotypic richness can be maintained at high levels if the life span of genets is long and sexual reproduction occurs occasionally (Eriksson 1993; Bengtsson 2003). The spatial arrangement of ramets in a genet (i.e. the clonal architecture) is known to affect mating opportunities and prospective intraspecific competition. Lophelia pertusa (henceforward referred to as Lophelia) is the main reef-building cold-water coral species in the NE Atlantic Ocean (Rogers 1999). It is gonochoristic; asexual reproduction occurs by fragmentation, and sexual reproduction follows an annual gametogenic cycle. The fecundity is very high: a 30-cm2 colony is estimated to have nearly 100,000 oocytes (Waller and Tyler 2005). After successful fertilization, larvae capable of spending up to 7 weeks in the water column develop (M. Dahl, pers. obs.). Reproduction by fragmentation allows clonal growth to act as a means of small-scale spatial dispersal by horizontal 123 Coral Reefs (2012) 31:1135–1148 spreading of clonal individuals (Gliddon et al. 1987). Hence, persistent clonal reproduction may enable genets (all genetically identical members of a clone) to grow laterally, and the age of individual Lophelia clones can be estimated by recording annual size increments (Witte and Stöcklin 2010). Some corals are capable of producing asexual propagules with long-distance dispersal abilities, which may help them to recover from environmental disturbances (van Oppen et al. 2008). To date, however, this reproductive strategy has not been reported in Lophelia. Lophelia occurs frequently along the European continental margin on ridges, seamounts and mound structures and in fjords (Rogers 1999) and shows broad-scale (1,000 km) genetic structuring (Le Goff-Vitry et al. 2004). The relative contribution of each mode of reproduction has been reported to vary among Lophelia populations along the continental margin (Le Goff-Vitry et al. 2004). However, these reported estimates of clonality might not be accurate since sampling was conducted with dredge and trawl where coral polyps may break and mix, making it hard to determine the precise sampling location of individual specimens. In contrast, recent advancements in sampling using camera-assisted ROVs now allow precise sampling, thereby allowing accurate assessment of clonal reproduction and mapping of the clonal architecture over entire reefs. In the North East Skagerrak, there are currently five known Lophelia reef localities. These reefs provide an excellent opportunity to map the distribution of individual clones and also assess the genetic structure among local reefs. In this study, we (1) assess the relative contribution of sexual and asexual reproduction, (2) assess the dispersal through clonal growth and larval transport and (3) discuss the implications for reef conservation. Materials and methods Study sites and sampling The Norwegian trench supplies the Skagerrak with Atlantic deep water, providing suitable habitats for cold-water corals. Through complex seabed topography, the Norwegian trench is further connected to the Oslofjord and the deep troughs running along the Swedish west coast. Polyps from a total of 142 Lophelia coral colonies were sampled from five spatially distinct coral reef complex localities in the NE Skagerrak (Fig. 1, Table 1). The locations span from the outer Oslofjord to the northern part of the Kosterfjord; the maximum distance between any reef pair is less than 35 km (Fig. 1). Coral samples were collected during numerous cruises over 6 years (2003–2009), preserved in ethanol (96 %) and maintained at -20 °C prior to Coral Reefs (2012) 31:1135–1148 1137 Table 1 Lophelia pertusa Ns G R Prop clones A Rare alleles Hexp Hobs FIS Fjellknausene 12 7 0.55 0.417 5.1 0.05 0.81 0.71 0.12 West Søstrene 4 4 1 0.000 5.8 0.10 0.88 0.75 0.15 East Søstrene 26 13 0.48 0.500 5.6 0.08 0.86 0.75 0.12* Tisler 87 35 0.40 0.598 5.6 0.17 0.84 0.72 0.15* Säcken 13 5 0.33 0.615 2.4 0.05 0.48 0.69 -0.44* Number of ramets (Ns), number of genets (G), genotypic richness (R), proportion of clones, allelic richness (A), rare alleles (expressed as mean number of alleles per locus normalized by genotypic richness), expected heterozygosity, observed heterozygosity and inbreeding coefficient (FIS) * significantly different from zero at p \ 0.05 genetic analyses. Samples were collected with a remotely operated vehicle (ROV) to minimize damage and to allow a precise geographical position of each sample. A detailed analysis of the spatial distribution of clones was performed for the Tisler reef (n = 87, Table 1), the largest known reef in the NE Skagerrak. Here, live coral extends over ca 250 ha (*20–30 % coverage, Fig. 2), and damage from extensive trawling activity has been documented (Lundälv 2003). Positions of the samples were obtained with a USBL underwater positioning system of type Simrad HPR 410 P in combination with a Furuno GPS gyro type SC 110 and a DGPS instrument type GBX Pro. Data were visualized and logged in the navigational software Olex. Accuracy of obtained positions was approximately ±2 m. All ramets belonging to a putative genet were sampled at a distance greater than the precision level of the positioning system. Genotyping DNA was extracted from coral polyp tissue with the Viogene Blood & Tissue Genomic DNA Extraction Miniprep System, following the manufacturer’s protocol. All samples were genotyped using thirteen microsatellite loci developed for Lophelia: three dinucleotide (Lp loci, LeGoff and Rogers 2002) and ten tri- or tetranucleotide loci (Lpe, Morrison et al. 2008). The Lp loci were amplified following LeGoff and Rogers (2002), whereas the Lpe loci were amplified in 10 lL PCR containing 2-60 ng/lL of template DNA, 0.05 U recombinant Taq DNA polymerase (TaKaRa TaqTM), 0.125 lM of forward and reverse primer, 1X buffer, 1.5 mM MgCl2 and 0.2 mM of each dNTP. PCR amplifications for Lpe loci were performed under the following conditions: initial denaturation at 94 °C (2 min), followed by 30 cycles at 94 °C (30 s), 58 °C (40 s) and 72 °C (30 s), with a final extension at 72 °C (10 min). Sets of three labelled primer pairs were poolplexed and sized on a CEQ 8000 Genetic Analysis System. Control for the presence of scoring errors due to stuttering during PCR amplification, null alleles and large allele dropout were performed with MICROCHECKER version 2.2.3 (van Oosterhout et al. 2004). Estimated frequencies of putative null alleles were subsequently calculated with FREENA (Chapuis and Estoup 2007) using the EM algorithm (Dempster et al. 1977) and the ENA method to calculate unbiased FST estimates adjusted for the presence of null alleles. Genotypic richness and diversity Estimation of components of clonal diversity (clonal richness, clonal heterogeneity and clonal evenness) was performed in GENCLONE version 2.0 (Arnaud-Haond and Belkhir 2007). Genotypic richness was determined as R = (G–1)/(N–1), where N is the number of genotyped samples and G the number of genotypes. GENCLONE was used to calculate the unique genotype probability (psex) to assess the presence of putative clonal genotypes as a result of sexual reproduction (Parks and Werth 1993). GENCLONE was also used to analyse intra-reef spatial genetic structure using spatial autocorrelation of genotypes (see Sokal and Oden 1978), represented by the kinship coefficient Fij (Loiselle et al. 1995). The spatial autocorrelation analysis was performed on two reef localities (Tisler and East Søstrene) where sample sizes allowed analysis of intra-reef spatial genetic structure. Spatial autocorrelation was also used to assess the impact of bottom trawling at Tisler reef by performing the analysis before and after removing all ramets sampled inside areas known to be affected by bottom trawling (Lundälv 2003). The spatial distance where the ramet and genet level correlograms intersect defines the clonal subrange (Harada et al. 1997; Alberto et al. 2005), a spatial measure that describes the scale at which clonality affects the genetic structure in a population. The clonal architecture of Tisler and East Søstrene was described with the spatial aggregation index (Ac) in GENCLONE. Finally, a permutation test was performed to obtain an accurate estimate of the minimum, average and maximum number of discriminated genetic individuals for the given number of samples from Tisler reef. 123 1138 Coral Reefs (2012) 31:1135–1148 West Søstrene (110 m) East Søstrene (95 - 120 m) Fjellknausene (95 - 115 m) Säcken (85 m) Tisler (70-145 m) Fig. 1 Sampling locations and depth for L. pertusa in north-east Skagerrak EstimateS (Colwell 2000) was used to calculate the expected total number of distinct genetic individuals at Tisler reef, by fitting an asymptotic function to the samplebased rarefaction curve from GENCLONE. The nonparametric richness estimators ICE and Chao2 were used (50 randomizations for each sample). Age of genets The minimum age of three different genets from Tisler reef was estimated by calculating the time required to grow linearly to cover the measured area. Linear growth rates for corals in this area have previously been measured to 5–7 mm year-1 (HERMES community report 2008). These three genets (referred to as the orange, blue and red genet) contained sufficient numbers of ramets to assume that the clone continuously covered the area. We used a conservative approach where only genets that appeared continuously distributed were included (see marked areas at the insert in Fig. 3). Individuals of the same clone scattered further afield and not within the continuous distribution were excluded 123 because ramets with a different genetic identity may confound the interpretation. During ROV sampling, no trawling damage was observed in the vicinity of these clones, something that would have caused an overestimation of the age estimate. The IVS FLEDERMAUS software was used to calculate the area covered by the clones. The calculated area was based on horizontal distances, thus representing horizontal coral growth. Since corals predominately grow vertically, age was calculated assuming an average coral colony height of 60 cm and a circular shape of the clone. By measuring a large number of coral colonies, we estimated an average 19-degree angle vertical growth. Repeated genotypes (ramets that belong to one genet) were included in the age and area coverage analysis described above. In subsequent genetic analysis, only unique multilocus genotypes were kept. Summary statistics Allele frequencies, observed heterozygosity (Ho), expected heterozygosity (He), inbreeding coefficients (f), Hardy– Coral Reefs (2012) 31:1135–1148 10° 56 59° 0.75 10° 56.5 1139 10° 57 10° 57.5 10° 58 10° 58.5 10° 59 10° 59.5 11° 0 59° 0.75 -240 -220 -200 -180 -160 -140 -120 -100 -80 -60 -40 -20 Water depth m 59° 0.5 59° 0.5 59° 0.25 59° 0.25 59° 0 59° 0 58° 59.75 58° 59.75 58° 59.5 58° 59.5 10° 56 10° 56.5 10° 57 10° 57.5 10° 58.5 10° 58 10° 59 10° 59.5 11° 0 Fig. 2 Bathymetric map of Tisler reef location generated from multibeam data. Reef is the dominant structure at the sill between the two basins. Photograph insert of Tisler reef illustrates the difficulty in distinguishing genetic individuals without molecular methods. The within-population demography is hidden by the identical appearance of different genetic individuals. The sponge Mycale lingua is commonly observed growing next to and competing with L. pertusa. Photograph by Tomas Lundälv Weinberg equilibrium (HWE) and genotypic linkage disequilibrium (LE) were calculated using GENEPOP 4.0.6 (Rousset 2008). Allelic richness using the rarefaction method was calculated using FSTAT version 2.9.3.2 (Goudet 2001). The significance levels were adjusted by Bonferroni correction when multiple tests were applied. statistic DK (Evanno et al. 2005) in the program STRUCTURE HARVESTER version 0.6.8 (Earl and vonHoldt 2011) to estimate the most likely number of K clusters. To corroborate the results from STRUCTURE and to assess recruitment and migration, we used genetic assignment performed with GENECLASS 2 (Piry et al. 2004), which is an individual-based classification method. Groups are defined a priori and individuals are assigned to known sources using the Bayesian allele frequency estimation method (Rannala and Mountain 1997) with the leave-oneout procedure. A simulation logarithm is implemented, which makes it possible to detect first-generation migrants (Paetkau et al. 2004). We performed the simulations with 10 000 individuals and a threshold p value of 0.01. Genetic differentiation Differentiation between sites was described by the FST estimator h (Weir and Cockerham 1984), and the null hypothesis of no differentiation was tested using Fisher’s exact test. We used factorial correspondence analysis implemented in GENETIX version 4.05.2 (Belkhir et al. 1996–2004) for visualizing the spatial variation in genetic composition among sampling localities. We used STRUCTURE 2.2 (Pritchard et al. 2000) to identify the number of different K genetic clusters. Five replicate runs were performed under the admixture model with correlated allele frequencies (K = 1–6; burn-in = 20,000 and 105 iterations). The program DISTRUCT (Rosenberg 2004) was used to plot individual membership assignments to each cluster. We calculated the ad hoc Results Genetic diversity All 13 microsatellite loci were highly polymorphic, averaging 22 alleles per locus and ranging from 7 (LpeC120) to 42 (Lp462), respectively (Electronic Supplemental 123 1140 Coral Reefs (2012) 31:1135–1148 Fig. 3 Lophelia pertusa. Bathymetric map of Tisler reef, facing westward, generated from multibeam data. Coloured dots represent spatial distribution of genets. Each colour is one genet; white dots indicate multilocus genotypes found only once. Insert is an enlargement of the central part of the reef. See Fig. 2 for scale Material, ESM Table S1); the mean number of alleles per population was 9.1. Locus-specific allele frequency distributions are shown in ESM Fig. S1, and locus-by-locus statistics on the number of alleles, Ho, He and FIS are presented in ESM Table S1. Overall loci values of He ranged between 0.48 and 0.88, Ho from 0.69 to 0.75 and FIS from -0.44 to 0.15 (Table 1). West Søstrene exhibited the highest allelic richness (5.8) in contrast to Säcken reef (2.4) that had values significantly lower than the average richness over all populations (4.88; p \ 0.001) (Table 1). Tisler had the highest mean number of private alleles (0.17), while Säcken reef and Fjellknausene had the lowest (0.05). Fisher’s exact test revealed that 16 of 65 locus/ population specific comparisons (24.6 %) deviated from Hardy–Weinberg expectations; six remained significant after Bonferroni correction (a = 0.05/65 = 0.0008). All deviations were due to heterozygote deficits, five were from Tisler reef and one from East Søstrene. Significant genotypic linkage disequilibrium was found in eight of 390 comparisons (2.1 %), but none remained significant after Bonferroni correction. Linkage between loci was found in three of 78 comparisons; only one of these comparisons (LpeD3 and Lp355) remained significant after Bonferroni correction. MICROCHECKER revealed no evidence of scoring errors, large allele dropouts or null alleles except in one case where the potential occurrence of null alleles in LpeC126 was indicated. The estimated frequency of the null allele varied between 0.27 and 0.39 in all populations, 123 and the unbiased FST values obtained using the ENA method did not vary significantly from the uncorrected ones. Overall uncorrected and corrected FST values across loci and populations were 0.038 (95 % CI: 0.029–0.047) and 0.039 (95 % CI: 0.030–0.047), respectively. Genotypic richness, indicating the relative importance of sexual versus asexual reproduction, ranged from 0.3 in the Säcken population to 1 in the West Søstrene population (Table 1). The probability that any of the different genets would have the same genotype by chance was significantly low (pgen \ 1.00E-15), and the probability that the repeated genotypes would have originated from distinct sexual reproductive events was also significantly low (psex \ 1.8E-24). Hence, all ramets with identical multilocus genotype were assumed to be clones. We found a total of 64 multilocus genotypes, and none of these were found at more than one sampling locality. This suggests that Lophelia does not develop asexual larvae and that coral fragments are not transported over long distances. Within-reef genetic structure The longest distance between two ramets of the same genet was 253 m (Fig. 3; black dots). In total, the 87 sampled ramets from Tisler reef were distributed among 35 genetic individuals (Fig. 4). The distribution of clonal ramets conformed to a Pareto distribution with a shallow slope Number of distinct MLG Coral Reefs (2012) 31:1135–1148 1141 34 32 30 28 26 24 22 20 18 16 14 12 10 8 6 4 2 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 44 46 48 50 52 54 56 58 60 62 64 66 68 70 72 74 76 78 80 82 84 86 88 Number of individuals Fig. 4 Lophelia pertusa. Boxplot showing the genotypic richness of Tisler reef samples. Central line depicts average number of genotypes. Edges of the box indicate minimum and maximum number of genotypes (b = 1.61) indicating low evenness. Clonal heterogeneity estimate and the equitability index (i.e. Simpson’s complement (D) = 0.94 and evenness (V) = 0.89, respectively) also show low evenness. Among the 35 genets detected at Tisler reef, fourteen appeared more than once (clones) and 21 were singletons. However, a few large clones dominated the reef; the three most frequently occurring genets made up 37 % of all sampled colonies. The most frequently sampled genet accounted for 16 % (Fig. 3; blue), while the second and third most frequently observed genets accounted for 13 and 8 %, respectively (Fig. 3; red and orange). Clonal distribution, heterogeneity and evenness estimates for East Søstrene were similar to those of Tisler (b = 1.95, D = 0.93, V = 0.83). East Søstrene had a higher evenness than Tisler, where six genets had ramets appearing more than once and seven were singlets. Indices of spatial aggregation were significant for both sites tested (p \ 0.001; Ac = 0.46 and 0.63, for Tisler and East Søstrene, respectively). The expected total number of genets at Tisler reef varied from a point estimate of 76 (ICE) to 90 (Chao2; 95 % confidence interval from 52 to 219). The estimated minimum age of genets at Tisler reef indicates that clones are of substantial age. The orange clone covered an area of 299 m2, which translates to an estimated age range of 4,408–6,172 years. Corresponding values for the blue and red clones are 164 and 130 m2, with estimated ages ranging from 3,251 to 4,569 and 2,906–4,068 years, respectively. Spatial autocorrelations revealed that the clonal subrange at Tisler reef extends up to 120 m. Ramets less than 28 m from each other have a 25 % probability of clonal identity. Probability of clonal identity decreased with increasing distances between ramets. When ramets sampled in trawled areas were excluded from the analysis, the clonal subrange decreased to 66 m (Fig. 5); the probability of clonal identity was higher at shorter distances (27 % at 25 m) and exhibited a steeper decrease with increasing distance (Fig. 5). The clonal subrange at East Søstrene extended 55 m. Additionally, the probability of clonal identity was high at short distances (57 % at 14 m) and decreased rapidly to 9 % at a distance of 60 m (data not shown). Population genetic structure The factorial correspondence analysis revealed that the four reef complexes located in the outer Oslofjord and outer Hvaler (West Søstrene, East Søstrene, Fjellknausene and Tisler) clustered on the right side of the first component axis, which explains most of the variation observed (41.3 %, Fig. 6). All genets from Säcken were distinct and formed a separate cluster in the multivariate space. The Bayesian clustering analysis in STRUCTURE (Fig. 7) and DK estimation (ESM Fig. S2 A, B) also suggested two distinct genetic clusters with Säcken reef as one cluster and the other four reefs as another. Likewise, individual assignment tests using GENECLASS showed that all individuals sampled at Säcken were assigned back to the Säcken location (Fig. 8), indicative of their distinct genetic composition. Fifty-five per cent of the individuals sampled at East Søstrene and 46 % of the individuals sampled at Tisler were assigned back to their sample location, whereas 36 % of individuals from East Søstrene were assigned to Tisler and 38 % of the individuals from Tisler were assigned to East Søstrene (Fig. 8). None of the individuals at Fjellknausene were assigned to its sampling location; all individuals were suggested to originate from East Søstrene (57 %) or Tisler (43 %). Similarly, at West 123 1142 Coral Reefs (2012) 31:1135–1148 Fig. 5 Spatial autocorrelation analysis of kinship coefficients for L. pertusa colonies at Tisler reef. Three different analyses are shown for both nontrawled (dashed lines) and trawled (filled lines) data sets in the diagram: (i) at ramet level (triangles); (ii) at genet level (circles); and (iii) the probability of clonal identity (squares) 0,3 Probability of clonal identity F(r) (87) Probability of clonal identity F(r) (82) Coancestry (Fij) All pairs of ramets included (87) Coancestry (Fij) All pairs of ramets included (82) Coancestry (Fij) Pairs of genets included (87) Coancestry (Fij) Pairs of genets included (82) 0,25 0,2 0,15 0,1 0,05 0 0 100 200 300 400 500 600 700 -0,05 Spatial distance (m) Axis 2 (22,49 %) 110 100 90 80 70 60 50 40 30 20 10 0 -10 -20 -30 -40 -50 -60 -70 -20 000 -18 000 -16 000 -14 000 -12 000 -10 000 8 000 4 000 -8 000 Axis 1 (41,31% ) -6 000 -4 000 -2 000 0 2 000 4 000 0 -4 000 -8 000 Axis -12 000 -16 000 3 (18,69 %) Fig. 6 Factorial correspondence analysis based on allele frequencies from 13 microsatellite loci genotyped in five L. pertusa populations from NE Skagerrak. Colours represent sample locations, orange = West Søstrene, green = East Søstrene, blue = Fjellknausene, yellow = Tisler, red = Säcken Søstrene, none were assigned back to that sampling location. Fifty per cent were assigned to Fjellknausene and 25 % to both Tisler and East Søstrene. Ten individuals across all populations were detected to be first-generation migrants. Of six migrants detected at Tisler reef, five originated from East Søstrene and one from West Søstrene. At East Søstrene, one first-generation migrant came from Tisler and a second from Fjellknausene. The two remaining first-generation migrants were found at Fjellknausene and West Søstrene; both originated from Tisler. Discussion 123 We present here the first empirical characterization of the fine-scale genetic structure of the cold-water coral L. pertusa. We show that reefs in Skagerrak consist of large and old clones, with an overall incidence of clonality of ca. 50 %. We found genetic differentiation and no sharing of clones among reefs, indicating that larval transport is the predominant mode of inter-reef dispersal. Coral Reefs (2012) 31:1135–1148 1143 Within-reef genetic structure and clonal spatial distribution K= 2 K= 3 K= 4 K= 5 en ck Sa ler Tis Fje l We lkna st use So n str e Ea en st e So str en e K= 6 Fig. 7 Bayesian STRUCTURE analysis of L. pertusa populations using combined data from 13 microsatellite loci. Results are shown for five levels of K (2–6). True K = 2 (DK = 65.3). Each individual is represented by a vertical line partitioned into K coloured segments that represent the individual’s estimated membership fractions. Black lines separate individuals from different sampling sites, which are labelled below the figure Genetic diversity was high in all Lophelia populations except at Säcken reef (Table 1). Säcken reef is characterized by low heterozygosity and allelic diversity, and such levels of diversity have been previously reported from East Africa coral populations (Souter and Grahn 2007; Ridgway et al. 2008). The other four reef populations exhibited levels of allelic richness similar to those described for many tropical, shallow-water coral species (e.g. Baums 2008; Shearer et al. 2009; Underwood et al. 2009). The total and mean number of alleles per locus for all reefs except Säcken reef were similar to average levels found in more than 70 populations of scleractinian corals (Shearer et al. 2009). Interestingly, the level of heterozygosity in Lophelia seems to be higher than those reported for many tropical coral species (cf. Table 3 in Baums 2008). This may be attributed to the gonochoristic breeding system that enforces outcrossing compared to the hermaphroditism common in tropical corals. However, these high levels of heterozygosity may be also expected in populations with high rates of clonal reproduction by retention of more alleles (Balloux et al. 2003). In addition, asexual reproduction and long generation times slow down the loss of genetic diversity through genetic drift (Orive 1993; Young et al. 1996). Despite the high values of expected heterozygosity in Lophelia, our results also show a considerable number of heterozygote deficiencies. Heterozygote deficiencies in adult sessile marine invertebrates with planktonic larvae Fig. 8 Individual-based selfassignment test using the leaveone-out procedure on five L. pertusa populations in NE Skagerrak. Vertical bars represent per cent individuals assigned back to sample location. Colours represent assigned locations: orange = West Søstrene, green = East Søstrene, blue = Fjellknausene, yellow = Tisler, red = Säcken 100% 75% 50% 25% Fjellknausene 0% West Søstrene East Søstrene Tisler reef Säcken reef 123 1144 have been reported in numerous studies (e.g. Johnson and Black 1982; Andrade and Solferini 2007; reviewed in Brownlow et al. 2008) and are also found in many studies of scleractinian corals (e.g. Ayre and Hughes 2000; van Oppen et al. 2008; Underwood et al. 2009). Sampling artefacts and attributes of molecular markers such as null alleles may contribute to these heterozygote deficiencies. However, clonal reproduction can largely account for this deficiency. Facultative clonal reproduction is common among sessile marine invertebrates (Hughes and Cancino 1985; Hughes 1989) and is also found in all hermatypic corals (Veron 2000). This life-history trait causes heterozygote deficiencies at the population level due to deviation from random mating (van Oppen et al. 2008). Clonal reproduction in Lophelia has been previously addressed in studies of genetic structure at larger spatial scales (LeGoff-Vitry and Rogers 2004; Morrison et al. 2011), where minimum and maximum distances between reefs ranged from 50 to 9,000 km. The proportions of clones were low (average 0.09 and 0.13, respectively), indicating that sexual reproduction is the predominant mode of reproduction. LeGoff-Vitry et al. (2004) reported evidence of clonality in only three out of ten populations investigated. La Galicia and Porcupine Seabight reefs had relatively low levels of asexual reproduction (0.15 and 0.20, respectively) compared to Darwin Mounds (0.49). The high proportion of clones at Darwin Mounds was attributed to low rates of sexual reproduction, patchy distribution of available habitat and bottom trawling. Similarly, Morrison et al. (2011) reported no clonality at 5 of 16 populations examined from mainly western Atlantic localities. The proportion of clones (0.42–0.62) and range of genotypic richness estimated in the present study (0.33–0.55, excluding the West Søstrene locality where only four specimens were collected) suggest that asexual reproduction in Lophelia is more important for reef development than previously shown. Thus, clonality may play a key role in cold-water coral reef establishment and maintenance. Our results suggest that asexual reproduction is the natural state found in cold-water coral reef development in the Skagerrak although trawling may also cause fragmentation. The dominance of few clones (as expressed by low evenness) has been observed in several clonal organisms (reviewed in Arnaud et al. 2007). Similarly, Lophelia reefs and shallow-water tropical corals appear dominated by few clones (Coffroth and Lasker 1998; Whitaker 2006; Baums et al. 2006). Intraspecific competition influences evenness, and this influence will have larger effects as population age increases and levels of sexual reproduction decrease. However, low evenness can also be explained by differences in time of establishment of genets (early established genets will have more time to increase their size and acquire resources). While a balance 123 Coral Reefs (2012) 31:1135–1148 between larval replenishment and genet longevity largely determines genotypic richness, evenness is more dependent on genet longevity and genet size and hence directly linked to clonal growth (Coffroth and Lasker 1998). Therefore, the evenness values calculated for Lophelia suggest that clonal growth contributes more to reef development than would be expected if only seedling recruitment were operating. Natural mechanisms of spatial extension of clones are governed by a relationship between growth and bioeroders such as clionid sponges that cause weakening and breakage of the skeleton. A growth pattern or patch development of Lophelia, often referred to as Wilson rings (Wilson 1979), arises from the initial settling of a larva and subsequent growth until the colony is so large and weakened that fragments fall off. These fragments will grow until they meet the same fate. The mechanism for the development of these rings supports the idea that clone size is directly related to genet age. Our estimates of genet ages suggest that the large clones are of considerable age, possibly being the same individuals that first settled after the glacial ice retreated several thousand years ago. Thus, genet longevity is an evolutionary consequence of clonal propagation and constitutes a key life-history trait of Lophelia. Consequently, the longevity and low number of genets observed in the present study suggest that the turnover rate is extremely slow. Unique genotypes (singletons) that appear within areas occupied by larger clones are indicative of subsequent immigration and settlement. Hence, development of populations is continuous with low rates of replenishment. Clonal reproduction also affects the reproductive dynamics within a population by influencing the clonal architecture and thus the opportunities for mating. Generally, there are two types of clonal architecture: (a) the ‘guerrilla’, characterized by a high level of intermingling of genotypes and (b) the ‘phalanx’, where high aggregations of clones result in a mosaic of clumped ramets of the same genotype. The phalanx architecture may decrease mate availability in gonochoristic species with external fertilization since the probability of successful fertilization decreases with distance. However, once this architecture is established, it can be advantageous for local persistence by optimizing resource capture and space occupation (Herben and Hara 1997). All genets at Tisler with one exception seem to conform to the phalanx strategy. The ‘blue’ clone exhibits a complex distribution of ramets that might have either a biological (fragmentation) or an anthropogenic (spreading by anchors) origin. Population structure The five Skagerrak reef complexes constitute two distinct genetic clusters: one comprised of the four reefs located in Coral Reefs (2012) 31:1135–1148 the outer Oslofjord and Hvaler area (West Søstrene, East Søstrene, Fjellknausene and Tisler) and a second consisting of only the isolated Säcken reef located further inside the coastal system of deep troughs. This pattern was supported by FST estimates (ESM Table S2), the high posterior probabilities in STRUCTURE (Fig. 7), as well as the individual assignment tests (Fig. 8). The gene flow/connectivity pattern observed within the larger cluster indicates asymmetric connectivity among subpopulations in a larger metapopulation. Further, the results suggest that Tisler and East Søstrene reefs are important sources of larvae. Such connectivity patterns are likely the result of the complex local hydrographic circulation conditions (Lavaleye et al. 2009), which in turn is attributed to the topographic complexity of the seabed in the area. In comparison to the restricted gene flow at a local scale within the NE Skagerrak, Lophelia populations on both sides of the Atlantic have been found to be both restricted and moderately connected genetically over larger spatial scales. LeGoff-Vitry et al. (2004) reported moderate levels of gene flow among populations on the European continental margin and isolated fjord populations. Recently, Morrison et al. (2011) showed significant population subdivision among Gulf of Mexico and West and East Atlantic Ocean, but high connectivity within regions. This geographical variation emphasizes that connectivity patterns are not solely a species-specific biological trait but also a reflection of local environmental conditions and stochastic oceanographic processes. Clonal reproduction influences population genetic structure in at least two ways: (1) reproductive output increases as a function of total genet size (Highsmith 1982; Hämmerli and Reusch 2003) and (2) the probability of genet death decreases as a function of number of ramets or size of genets/ramets (Highsmith 1982). In the coral Goniastrea aspera, for example, the largest colonies produce 25 % of the annual egg production even though numerically these colonies only comprise 3 % of the population (Babcock 1984). Hence, large genets, as was observed at Tisler reef, may contribute more to the gametic gene pool. Once genets become dominant, the probability of genet death approaches zero and their genes will persist in the population until the ancient genotype dies. This ‘swamping’ of the local gene pool means that most larvae are descendants of the dominant genotype at a reef. Implication for conservation and future directions A striking feature of cold-water coral reefs is that they are generally composed of single coral species. In tropical coral reef ecosystems, community composition (species, species diversity and abundance) likely affects the ability to respond to environmental changes (Connell et al. 2004). 1145 In contrast, cold-water coral reefs such as Lophelia appear to be more dependent on the genotypic diversity within populations to respond to environmental changes. Recently, it has been recognized that genetic diversity may influence ecological processes at all possible levels of organization (Hughes et al. 2008). Genetic diversity has been shown to have similar effects on, for example, fitness, responses to disturbance and ecosystem function as species diversity has for a wide range of organisms including plants, invertebrates and vertebrates (Gamfeldt et al. 2005; Pearman and Garner 2005; Johnson et al. 2006; Crutsinger et al. 2006; Mattila and Seeley 2007; Hughes and Stachowicz 2009). Clonal species with high genotypic diversity are described as having higher resistance to parasites and pathogens (Booth and Grime 2003; Altermatt and Ebert 2008). Additionally, experimental studies have shown positive effects of genotypic diversity on survival and faster recovery after extreme climatic events (Hughes and Stachowicz 2004; Reusch et al. 2005). Here, we show that the genotypic diversity was typically less than 50 % for the reefs. We detected 35 genetically distinct individuals representing at best 90 individuals in total at the largest reef, Tisler. Shearer et al. (2009) estimated from tropical corals that 35 randomly sampled colonies are required to maintain [90 % of the genetic diversity. Taking these estimates as reference, for highly clonal species as Lophelia, nearly 100 colonies would be needed to maintain similar levels of diversity. Rare species have been prioritized in conservation efforts due to the high risk of extinction, but it has become increasingly clear that more common habitat-forming species such as Lophelia may be equally important to conserve (Gaston and Fuller 2008). Conservation efforts in Lophelia could be improved with information of key life-history traits such as levels of clonality and genetic diversity within populations. For example, results from this study on clumped spatial distribution of clones indicate that trawling activities can eliminate unique genets from the population. At Tisler, there are large areas of dead coral structure at both ends of the reef surrounded by numerous trawl scars (Lundälv 2003). Our results show an increase in the clonal subrange at Tisler reef where high trawling activity has been documented (Fig. 5). The large distances between ramets suggest that bottom trawling has altered the genetic structure. Prior to these damages, the reef had a length of about 2 km (Lavaleye et al. 2009); at present, the length of living coral reef is approximately 1,200 m. Trawling on cold-water coral habitats disrupts the threedimensional structure of reefs and subsequently alters the hydrodynamic and sedimentary conditions around it (Rogers 1999). Genetic changes in the form of increased clonal subranges have also been observed in seagrass habitats in the Mediterranean Sea (Diaz-Almela et al. 123 1146 2007). However, to our knowledge, changes in the genetic structure of scleractinian corals specifically produced by trawling have not been previously reported. Thus, description of genetic and genotypic diversity distribution is increasingly urgent for cold-water coral ecosystems. Due to longevity of individuals in cold-water corals, even small genetic differences between Lophelia populations should be considered in conservation decisions. Sampling strategy, accuracy and spatial resolution will all affect the measured level of genotypic richness and should therefore be stated explicitly. Description of the genetic diversity distribution is an important step, but the factors and processes responsible for the observed patterns remain untested. In 2009, Sweden and Norway established two protected areas, which together form a transboundary marine national park covering 751 km2. The park encompasses all the reefs examined in the present study; thus, all reefs are protected. The Säcken reef is in a severely degraded condition; the low genetic variation in combination with the apparent lack of gene flow to Säcken reef from other reefs suggests that the potential for future adaptation to environmental change may be low. The distances between the living colonies also suggest that successful fertilization may be limited. Consequently, active restoration of Säcken reef population is recommended. The present study demonstrates the usefulness of investigating the small-scale spatial genetic structure to improve our understanding of reef development, maintenance and the conservation of these habitats. Acknowledgments We thank Lisbeth Jonsson and Armin Form for assistance in collecting samples and Sophie Arnaud-Haond for valuable advice and comments on an earlier version. Funding was provided by grants from Oscar and Lili Lamm, Helge Ax:son Johnsons stiftelse, Wilhelm och Martina Lundgrens Vetenskaps- och Understödsfond and Colliander. We would also like to acknowledge the help of three anonymous reviewers whose suggestions helped to improve the manuscript considerably. Funding for sampling was supported by the two European projects: HERMES EC contract no GOCE-CT-2005-511234, funded by the European Commission’s Sixth Framework Programme under the priority ‘Sustainable Development, Global Change and Ecosystems’ and the HERMIONE project, EC contract no 226354, funded by the European Commission’s 7th Framework Programme under the theme Environment (including climate change). References Alberto F, Gouveia L, Arnaud-Haond S, Pérez-Lloréns J, Duarte CM, Serrao EA (2005) Within-population spatial genetic structure, neighbourhood size and clonal subrange in the seagrass Cymodocea nodosa. Mol Ecol 14:2669–2681 Altermatt F, Ebert D (2008) Genetic diversity of Daphnia magna populations enhances resistance to parasites. Ecol Lett 11:1–11 Andrade SC, Solferini VN (2007) Fine-scale genetic structure overrides macro-scale structure in a marine snail: nonrandom 123 Coral Reefs (2012) 31:1135–1148 recruitment, demographic events or selection? Biol J Linn Soc 91:23–36 Arnaud-Haond S, Belkhir K (2007) GENCLONE: a computer program to analyse genotypic data, test for clonality and describe spatial clonal organization. Mol Ecol Notes 7:15–17 Ayre DJ, Hughes TP (2000) Genotypic diversity and gene flow in brooding and spawning corals along the Great Barrier Reef, Australia. Evolution 590:1590–1605 Babcock RC (1984) Reproduction and distribution of two species of Goniastrea (Scleractinia) from the Great Barrier Reef Province. Coral Reefs 2:187–195 Balloux F, Lehmann L, de Meeûs T (2003) The population genetics of clonal and partially clonal diploids. Genetics 164:1635–1644 Baums IB (2008) A restoration genetics guide for coral reef conservation. Mol Ecol 17:2796–2811 Baums IB, Miller MW, Hellberg ME (2006) Geographic variation of clonal structure in a reef-building Caribbean coral, Acropora palmata. Ecol Monogr 76:503–519 Belkhir K, Borsa P, Chikhi L, Raufaste N, Bonhomme F (1996–2004) GENETIX 4.05, logiciel sous Windows TM pour la génétique des populations. Laboratoire Génome, Populations, Interactions, CNRS UMR 5171, Université de Montpellier II, Montpellier (France) Bengtsson BO (2003) Genetic variation in organisms with sexual and asexual reproduction. J Evol Biol 16:189–199 Booth RE, Grime JP (2003) Effects of genetic impoverishment on plant community diversity. J Ecol 91:721–730 Brownlow RJ, Dawson DA, Horsburgh GJ, Bell JJ, Fish JD (2008) A method for genotype validation and primer assessment in heterozygote-deficient species, as demonstrated in the prosobranch mollusc Hydrobia ulvae. BMC Genet 9:55 Chapuis MP, Estoup A (2007) Microsatellite null alleles and estimation of population differentiation. Mol Biol Evol 24:621– 631 Coffroth MA, Lasker HR (1998) Population structure of a clonal gorgonian coral: the interplay between clonal reproduction and disturbance. Evolution 52:379–393 Colwell RK (2000) EstimateS version 8.2.0: statistical estimation of species richness and shared species from samples. Freeware published at http://viceroy.eeb.uconn.edu/EstimateS Connell JH, Hughes TP, Wallace CC, Tanner JE, Harms KE, Kerr AM (2004) A long-term study of competition and diversity of corals. Ecol Monogr 74:179–210 Cowen RK, Lwiza KMM, Sponaugle S, Paris CB, Olson DB (2000) Connectivity of marine populations: open or closed? Science 287:857–859 Crutsinger GM, Collins MD, Fordyce JA, Gompert Z, Nice CC, Sanders NJ (2006) Plant genotypic diversity predicts community structure and governs an ecosystem process. Science 313:966– 968 de Meeûs T, Prugnolle F, Agnew P (2007) Asexual reproduction: genetics and evolutionary aspects. Cell Mol Life Sci 64:1355– 1372 Dempster AP, Laird NM, Rubin DB (1977) Maximum likelihood from incomplete data via the EM algorithm. J R Stat Soc B 39: 1–38 Diaz-Almela E, Arnaud-Haond S, Vliet MS, Álvarez E, Marbà N, Duarte CM, Serrão EA (2007) Feed-backs between genetic structure and pertubation-driven decline in seagrass (Posidonia oceanica) meadows. Conserv Genet 8:1377–1391 Dorken ME, Eckert CG (2001) Severely reduced sexual reproduction in northern populations of a clonal plant, Decodon verticillatus (Lythraceae). J Ecol 89:339–350 Earl DA, vonHoldt BM (2011) STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and Coral Reefs (2012) 31:1135–1148 implementing the Evanno method. Conservation Genet Resour. doi:10.1007/s12686-011-9548-7 Eckert CG (2002) The loss of sex in clonal plants. Evol Ecol 15:501–520 Eriksson O (1989) Seedling dynamics and life histories in clonal plants. Oikos 55:231–238 Eriksson O (1993) Dynamics of genets in clonal plants. Trends Ecol Evol 8:313–316 Evanno G, Regnaut S, Goudet J (2005) Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Mol Ecol 14:2611–2620 Gamfeldt L, Wallen J, Jonsson PR, Berntsson KM, Havenhand JN (2005) Increasing intraspecific diversity enhances settling success in a marine invertebrate. Ecology 86:3219–3224 Gaston KJ, Fuller RA (2008) Commonness, population depletion and conservation biology. Trends Ecol Evol 23:14–19 Gladychev EA, Meselson M, Arkhipova I (2008) Massive horizontal gene transfer in Bdelloid rotifers. Science 320:1210–1213 Gliddon C, Belhassen E, Gouyon PH (1987) Genetic neighborhoods in plants with diverse systems of mating and different patterns of growth. Heredity 59:29–32 Goudet (2001) FSTAT, a program to estimate and test gene diversities and fixation indices (version 2.9.3) Hämmerli A, Reusch TBH (2003) Genetic neighbourhood of clone structures in eelgrass meadows quantified by spatial autocorrelation of microsatellite markers. Heredity 91:448–455 Harada Y, Kawano S, Iwasa Y (1997) Probability of clonal identity: inferring the relative success of sexual versus clonal reproduction from spatial genetic patterns. J Ecol 85:591–600 Harper JL (1977) Population biology of plants. Academic Press, London Hartnett DC, Bazzaz FA (1985) The genet and ramet population dynamics of Solidago canadensis in an abandoned field. J Ecol 73:407–413 Hauser L, Carvalho GR (2008) Paradigm shifts in marine fisheries genetics: ugly hypotheses slain by beautiful facts. Fish Fish 9: 333–362 Herben T, Hara T (1997) Competition and spatial dynamics of clonal plants. In: de Kroon H, van Groenendael J (eds) The ecology and evolution of clonal plants. Backhuys Publisher, Leiden, pp 331–357 HERMES (2008) Periodic activity report: month 36. Hotspot ecosystem research on the margins on the European seas. FP6 integrated project sustainable development, global change and ecosystems. EC contract No. GOCE-CT-2005-511234 Highsmith RC (1982) Reproduction by fragmentation in corals. Mar Ecol-Prog Ser 7:207–226 Hoffmann RJ (1986) Variation in contributions of asexual reproduction to the genetic structure of populations of the sea anemone Metridium senile. Evolution 40:357–365 Honnay O, Bossuyt B (2005) Prolonged clonal growth: escape route or route to extinction? Oikos 108:427–432 Hughes RN (1989) A functional biology of clonal animals. Chapman & Hall, London and New York Hughes RN, Cancino JM (1985) An ecological overview of cloning in metazoa. In: Jackson JBC, Buss LW, Cook RE (eds) Population biology and evolution of clonal organisms. Yale University Press, New Haven Hughes AR, Stachowicz JJ (2004) Genetic diversity enhances the resistance of a seagrass ecosystem to disturbance. Proc Natl Acad Sci USA 101:8998–9002 Hughes AR, Stachowicz JJ (2009) Ecological impacts of genotypic diversity in the clonal seagrass Zostera marina. Ecology 90: 1412–1419 Hughes AR, Inouye BD, Johnson MTJ, Underwood N, Vellend M (2008) Ecological consequences of genetic diversity. Ecol Lett 11:609–623 1147 Johnson MS, Black R (1982) Chaotic genetic patchiness in an intertidal limpet, Siphonaria sp. Mar Biol 70:157–164 Johnson MTJ, Lajeunesse MJ, Agrawal AA (2006) Additive and interactive effects of plant genotypic diversity on arthropod communities and plant fitness. Ecol Lett 9:24–34 Lasker HR, Coffroth MA (1999) Responses of clonal reef taxa to environmental change. Am Zool 39:92–103 Lavaleye M, Duineveld G, Lundälv T, White M, Guihen D, Kiriakoulakis K, Wolff G (2009) Cold-water corals on the Tisler reef: preliminary observations on the dynamic reef environment. Oceanography 22:76–84 Le Goff MC, Rogers AD (2002) Characterization of 10 microsatellite loci for the deep-sea coral Lophelia pertusa (Linnaeus 1758). Mol Ecol Notes 2:164–166 Le Goff-Vitry MC, Pybus OG, Rogers AD (2004) Genetic structure of the deep-sea coral Lophelia pertusa in the northeast Atlantic revealed by microsatellites and internal transcribed spacer sequences. Mol Ecol 13:537–549 Lesica P, Allendorf FW (1995) When are peripheral populations valuable for conservation? Conserv Biol 9:753–760 Loiselle BA, Sork VL, Nason J, Graham C (1995) Spatial genetic structure of a tropical understory shrub, Psychotria officinalis (Rubiaceae). Am J Bot 82:1420–1425 Loveless MD, Hamrick JL (1984) Ecological determinants of genetic structure in plant populations. Annu Rev Ecol Syst 15:65–95 Lundälv T (2003) Kartläggning av marina habitat i Yttre Hvaler, nordöstra Skagerrak. En pilotstudie. Rapport till Fylkesmannen i Østfold och Nordiska Ministerrådet, p 16 Mattila HR, Seeley TD (2007) Genetic diversity in honey bee colonies enhances productivity and fitness. Science 317:362–364 Morrison CL, Eackles MS, Johnson RL, King TL (2008) Characterization of 13 microsatellite loci for the deep-sea coral, Lophelia pertusa (Linnaeus 1758), from the western North Atlantic Ocean and Gulf of Mexico. Mol Ecol Resour 8:1037–1039 Morrison CL, Ross SW, Nizinski MS, Brooke S, Järnegren J, Waller RG, Johnson RL, King TL (2011) Genetic discontinuity among regional populations of Lophelia pertusa in the North Atlantic Ocean. Conserv Genet 12:713–729 Orive ME (1993) Effective population-size in organisms with complex life-histories. Theor Popul Biol 44:316–340 Paetkau D, Slade R, Burden M, Estoup A (2004) Genetic assignment methods for the direct, real-time estimation of migration rate: a simulation-based exploration of accuracy and power. Mol Ecol 13:55–65 Pan JJ, Price JS (2002) Fitness and evolution in clonal plants: the impact of clonal growth. Evol Ecol 15:583–600 Parks JC, Werth CR (1993) A study of spatial features of clones in a population of bracken fern, Pteridium aquilinum (Dennstaedtiaceae). Am J Bot 80:537–544 Pearman PB, Garner TWJ (2005) Susceptibility of Italian agile frog populations to an emerging strain of Ranavirus parallels population genetic diversity. Ecol Lett 8:401–408 Piry S, Alapetite A, Cornuet JM, Paerkau D, Baudouin L, Estoup A (2004) GENECLASS2: a software for genetic assignment and first-generation migrant detection. J Hered 95:536–539 Pritchard JK, Stephens M, Donnelly P (2000) Inference of population structure using multilocus genotype data. Genetics 155:945–959 Rannala B, Mountain JL (1997) Detecting immigration by using multilocus genotypes. Proc Natl Acad Sci USA 94:9197–9201 Reusch TBH, Ehlers A, Hammerli A, Worm B (2005) Ecosystem recovery after climatic extremes enhanced by genotypic diversity. Proc Natl Acad Sci USA 102:2826–2831 Ridgway T, Riginos C, Davis J, Hoegh-Guldberg O (2008) Genetic connectivity patterns of Pocillopora verrucosa in southern African Marine Protected Areas. Mar Ecol-Prog Ser 354:161– 168 123 1148 Rogers AD (1999) The biology of Lophelia pertusa (Linnaeus 1758) and other deep-water reef-forming corals and impacts from human activities. Int Rev Hydrobiol 84:315–406 Rosenberg NA (2004) DISTRUCT: a program for the graphical display of population structure. Mol Ecol Notes 4:137–138 Rousset F (2008) GENEPOP’ 007: a complete re-implementation of the GENEPOP software for Windows and Linux. Mol Ecol Resour 8:103–106 Shearer TL, Porto I, Zubillaga AL (2009) Restoration of coral populations in light of genetic diversity estimates. Coral Reefs 28:728–733 Silvertown J (2008) The evolutionary maintenance of sexual reproduction: evidence from the ecological distribution of asexual reproduction in clonal plants. Int J Plant Sci 169:157–168 Sokal R, Oden N (1978) Spatial autocorrelation in biology 2. Some biological implications and four applications of evolutionary and ecological interest. Biol J Linn Soc 10:229–249 Souter P, Grahn M (2007) Spatial genetic patterns in lagoonal, reefslope and island populations of the coral Platygyra daedalea in Kenya and Tanzania. Coral Reefs 27:433–439 Takabayashi M, Carter DA, Lopez JV, Hoegh-Guldberg O (2003) Genetic variation of the scleractinian coral Stylophora pistillata, from western Pacific reefs. Coral Reefs 22:17–22 Underwood JN, Smith LD, van Oppen MJH, Gilmour JP (2009) Ecologically relevant dispersal of corals on isolated reefs: implications for managing resilience. Ecol Appl 19:18–29 van Kleunen M, Fisher M, Schmid B (2001) Effects of intraspecific competition on size variation and reproductive allocation in a clonal plant. Oikos 94:5151–5524 123 Coral Reefs (2012) 31:1135–1148 van Oosterhout C, Hutchinson WF, Willis DPM, Shipley P (2004) MICRO-CHECKER: a software for identifying and correcting genotyping errors in microsatellite data. Mol Ecol Resour 4: 535–538 van Oppen MJH, Gates RD (2006) Conservation genetics and the resilience of reef-building corals. Mol Ecol 15:3863–3883 van Oppen MJH, Lutz AH, De0 ath AG, Peplow L, Kininmonth SJ (2008) Genetic traces of recent long-distance dispersal in a predominantly self-recruiting coral. PLoS One 3(10):e3401 Veron JEN (2000) Corals of the world. Australian Institute of Marine Science, Townsville Waller RG, Taylor PA (2005) The reproductive biology of two deepwater, reef building scleractinians from the NE Atlantic Ocean. Coral Reefs 24:514–522 Weir B, Cockerham CC (1984) Estimating F-statistics for the analysis of population structure. Evolution 38:1358–1370 Whitaker K (2006) Genetic evidence for mixed modes of reproduction in the coral Pocillopora damicornis and its effect on population structure. Mar Ecol-Prog Ser 306:115–124 Wilson JB (1979) ‘‘Patch’’ development of the deep-water coral Lophelia pertusa (L) on Rockall Bank. J Mar Biol Assoc UK 59:165–177 Witte LC, Stöcklin J (2010) Longevity of clonal plants: why it matters and how to measure it. Ann Bot 106:859–870 Young A, Boyle T, Brown T (1996) The population genetic consequences of habitat fragmentation for plants. Trends Ecol Evol 11:413–418