You are on page 1of 11

Biological Journal of the Linnean Society, 2008, 95, 106116.

With 4 gures

Is life history a barrier to dispersal? Contrasting patterns of genetic differentiation along an oceanographically complex coast
CRAIG D. H. SHERMAN*, ALISON HUNT and DAVID J. AYRE
Institute for Conservation Biology, School of Biological Sciences, The University of Wollongong, Wollongong, New South Wales 2522, Australia
Received 10 September 2007; accepted for publication 18 December 2007

Extreme variation in early life-history strategies is considered a moderately good predictor of genetic subdivision and hence dispersal for a range of marine species. In reality, however, a good deal of population differentiation must reect historical effects, more subtle variation in life histories, and, particularly, the interaction of larvae with oceanographic processes. Using a combination of allozyme and microsatellite markers, we show that the large-scale genetic structure of populations of three species (direct and planktonically developing cushion stars and a planktonic developing sea anemone that is also asexually viviparous) varies consistently, in line with the predicted capacity for dispersal within three geographic regions. We detected high levels of genetic subdivision for the direct developing cushion star (FST = 0.6), low levels for the planktonically developing cushion star (FST = 0.009), and intermediate levels for the sexual/asexual sea anmone (FST = 0.19). These patterns are exhibited despite the highly variable patterns of current movement and the presence of biogeographic barriers. Our results suggest that, although there is large scale genetic differentiation for two species, patterns of population connectivity are remarkably consistent within major regions and do not reect variation in major oceanographic processes or genetic discontinuity coincident with biogeographic boundaries. 2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116.

ADDITIONAL KEYWORDS: gene ow intertidal invertebrates larval dispersal.

INTRODUCTION
The early life histories of marine organisms can vary dramatically, ranging from species with direct development and highly philopatric dispersal through to species with planktotrophic larvae that spend months in the water column and have the potential for passive dispersal over large distances (Strathmann, 1985; Palumbi, 1994). However, predicting the extent to which dispersal will inuence population structure is confounded both by the extent of life history variation and the interaction of larvae with both the biotic and abiotic environments, and especially oceanographic processes. At the extreme, poor dispersers should have highly subdivided populations that are often individually lacking in genetic variation as a
*Corresponding author. E-mail: csherman@uow.edu.au

result of founder events and subsequent genetic drift. Although the effects of drift and founder events may be most severe for populations at the species range limit, dispersal among populations anywhere within the species range will still require rafting, with consequent genetic subdivision and reduced levels of genetic diversity. By contrast, good dispersers should maintain large and relatively homogeneous populations with little or no genetic subdivision. Moreover, a meta analysis by Bohonak (1999) suggests that estimates of the dispersal ability of marine, freshwater, and terrestrial taxa are a moderately good predictor of levels of genetic subdivision (r2 = 0.52). However, this relationship is clearly complex, especially for marine organisms where the scale and pattern of larval dispersal is likely to be strongly inuenced by the magnitude and direction of prevailing currents and ne-scale topographical variation (Hedgecock,

106

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

CONTRASTING PATTERNS OF GENETIC DIFFERENTIATION 1986; Williams & Benzie, 1993; Palumbi, 1994; Lessios, Kessing & Pearse, 2001; Ayre & Hughes, 2004). In the present study, we use a hierarchical survey of genetic variation along 2500 km of the south east cost of Australia to test not only whether variation in early life history strategies is a good predictor of genetic subdivision and hence the dispersal potential of larvae, but also to investigate how genetic subdivision varies under changing oceanographic conditions. This region includes two major but overlapping biogeographic provinces: the Peronian Province, extending from northern New South Wales (NSW) into eastern Victoria and northern Tasmania, and the Maugean Province, encompassing Tasmania and eastern Victoria (Bennett & Pope, 1953). However, we argue that, from the perspective of larval transport, it is reasonable to divide the more northerly Peronian Province into two regions reecting a well documented change in the reliability of the Eastern Australian current (Murray-Jones & Ayre, 1997). Our target species were the cushion stars Parvulastra exigua (= Patiriella exigua; Lamarck, 1816) and Meridiastra calcar (= Patiriella calcar; Lamarck, 1816), and the asexually viviparous sea anemone Actinia tenebrosa (Farquhar, 1898). They are representative of three contrasting life histories and so may be expected to display contrasting genetic population structures either across their ranges or among geographic regions. Paradoxically, however, all three species have similarly south-eastern Australian geographic ranges occupying almost every suitable rocky headland across more than 1500 km (although M. calcar and A. tenebrosa have more extensive northern ranges). Meridiastra calcar and P. exigua provide the sharpest life history contrasts. Meridiastra calcar is expected to be a good disperser because it produces lecithotrophic planktonic larvae that hatch from benthic egg masses, whereas P. exigua is a direct developer, hatching from benthic eggs with no planktonic stage (Lawson-Kerr & Anderson, 1978; Byrne, 1992). Actinia tenebrosa is potentially a good disperser. Its populations are maintained principally by localized asexual recruitment (Ayre, 1983, 1984; Sherman, Peucker & Ayre, 2007); however, it also displays an annual cycle of gonad maturation (Sherman et al., 2007) and genetic data suggest that populations are founded by planktonically dispersed sexually produced larvae (Ayre, 1984; Ayre, Read & Wishart, 1991). Nevertheless, the successful dispersal of sexually generated larvae is expected to be a rare event for A. tenebrosa because only three mature females have been detected in dissections of hundreds of individuals over multiple years (Sherman et al., 2007; D. J. Ayre, unpubl. data). We therefore expect

107

Figure 1. The location of collection sites for Parvulastra exigua (prev. Patiriella exigua), Meridiastra calcar (prev. Patiriella calcar), and Actinia tenebrosa along the east coast of Australia. EAC, East Australian Current.

the combination of clonal expansion and episodic dispersal to produce considerable population differentiation. Limited population genetic data are available for each of these species supporting these basic contrasts. For P. exigua and M. calcar, this is largely limited to one geographic region (Hunt, 1993) and, for A. tenebrosa, the available data span a broader range but the interpretation of genetic structure was problematic because the evidence for an apparent restriction of gene ow between the central and southern regions was largely based on clinal variation at a pair of tightly linked loci (Ayre et al., 1991). Additional detailed and spatially extensive genetic surveys are therefore needed to examine the link between life history and genetic differentiation along this complex shore.

MATERIAL AND METHODS COLLECTION OF SPECIMENS


We collected samples of each of the three study species from two to three geographic regions, covering up to 2580 km, along the east coast of Australia (Fig. 1). We selected regions on the basis of current regimes along the Australian east coast. The East Australian Current (EAC) ows southward along the east Australian coast, showing remarkable levels of latitudinal and seasonal variation, but is strongest between 25S and 32S during the summer periods. On the northern and central NSW coast (region 1), larvae should therefore experience relatively reliable southward transport. Near Sydney, the current begins to separate from the coast and move off-shore into the Tasman Sea (Chiswell, Toole & Church, 1997; Ridgway & Godfrey, 1997). Episodes of strong south-

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

108

C. D. H. SHERMAN ET AL. MDH; 6-phosphogluconate dehydrogenase (6PGD, E.C. 1.1.1.44) and octopine dehydrogenase (ODH, E.C. 1.5.1.11). In addition to the four alloyme loci, we genotyped all A. tenebrosa samples for four microsatellite loci At1, At5, At21a, and At38 using the methodology of Sherman et al. (2007).

ward ow are a typical summer feature throughout the NSW south coast (region 2). This region is bounded to the south by the 145 km extent of the Ninety Mile Beach (broken only by one tiny natural rock platform), thus representing a physical barrier without the hard substrate normally required to support many intertidal taxa. Community level surveys of biodiversity imply that a biographic barrier exists in this area (Hidas et al., 2007). At the southern tip of eastern Australia, the weaker EAC passes the eastern entrance of Bass Straight where some of the water is deected into the Bass Straight and mixes with cooler South Australian Current water. The rest of the EAC continues towards Tasmania where it ows down its east coast. We follow Bennett & Pope (1953) in treating this relatively complex area as a single third region, although we recognize that arguments could be made for further subdivision. We therefore predict that successful larval dispersal, and hence gene ow, may become progressively less reliable from northern to southern NSW and that gene ow from and to the south may potentially be disrupted by Bass Strait and the Ninety Mile Beach biogeographic barrier. For the direct developer P. exigua, we collected 3348 individuals from 20 populations covering the central and southern regions. For the broadcast spawner M. calcar, we collected 2749 individuals from each of 26 populations covering all three regions. For the broadcast spawner/asexual brooder A. tenebrosa, we collected 24 individuals from each of 19 populations covering all three geographic regions. As estimates of inter-population genetic variation may be biased by clonal replication (Ayre, 1984; Stoddart, 1988), we attempted to minimize the collection of clonemates of A. tenebrosa by taking only widely spaced individuals within each population. All samples were transported to the laboratory alive and then stored frozen at -80 C pending electrophoretic and DNA analysis.

STATISTICAL

ANALYSIS

ALLOZYME

ELECTROPHORESIS AND

DNA

ANALYSIS

The methodology for allozyme electrophoresis of M. calcar and P. exigua is provided in Hunt (1993) and for A. tenebrosa in Ayre (1982). For M. calcar, each specimen was assayed for ve enzyme enconding loci: glucose-phosphate isomerase (GPI, EC 5.3.1.9); phosphoglucomutase (PGM, EC 5.4.2.2); hexokinase (HK, EC 2.7.1.1); malate dehydrogenase (MDH, EC 1.1.1.37); and superoxide dismutase (SOD, EC 1.15.1.1). For P. exigua, each specimen was assayed for six loci; GPI, PGM, HK, MDH, and L-leucyl-L-tyrosine peptidase (LTP, EC 3.4.11.1) and L-leucyl-L-proline peptidase (LPP, EC 3.4.11.1). For A. tenebrosa, the enzymes assayed were GPI;

To determine whether each locus assorted independently we tested each pairwise combination of loci for linkage disequilibrium (Weir, 1979) within each population using the software GENEPOP, version 3.4 (Raymond & Rousset, 1995) and applied a sequential Bonferroni correction to correct for multiple tests (Rice, 1989). We found no signicant association among loci for P. exigua or M. calcar. For A. tenebrosa, we detected 84 signicant associations out of 454 possible comparisons, with only 19 remaining signicant after the application of a sequential Bonferroni correction. However, as clonal reproduction can generate nonrandom associations among loci and may mimic the affects of physical linkage over the entire genome, reanalysis using only unique multi-locus genotypes was carried out. This analysis revealed only 19 signicant associations out of 454 possible comparisons, with none remaining signicant after the application of a sequential Bonferroni correction. We used a number of analyses to characterize the effect of life history variation on patterns of genetic differentiation and connectivity among populations and regions. All analyses were carried using the genetics program ARLEQUIN, version 3.01 (Excoffier, Laval & Schneider, 2005) unless otherwise indicated. We examined population genetic structure using a hierarchical analysis of molecular variance (AMOVA; Excoffier, Smouse & Quattro, 1992) and standardized F-statistics (Weir & Cockerham, 1984). Genetic variation was partitioned among regions (FRT), populations within regions (FSR), and among all populations (FST). As asexual reproduction can inuence estimates of population structure (Ayre & Hughes, 2000; Billingham et al., 2003; Baums, Miller & Hellberg, 2006), we assessed the importance of asexual reproduction within populations of A. tenebrosa by rst comparing the number of colonies sampled (N) to the number of unique multi-locus genotypes (Ng) detected. We then compared the ratio of observed multi-locus genotypic diversity (Go) to that expected under conditions of sexual reproduction with free recombination (Ge), as described by Stoddart & Taylor (1988). The statistical signicance of differences between Go and Ge was assessed by determining whether Go lay outside the 95% condence interval of Ge (Stoddart & Taylor, 1988). To reduce the chance of type I errors, a sequential Bonferroni correction was then applied. Variation

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

CONTRASTING PATTERNS OF GENETIC DIFFERENTIATION in levels of genotypic diversity among regions were assessed using one-way analysis of variance (ANOVA) with all variables meeting the assumptions of normality (Anderson Darling test statistic, P > 0.11) and homogeneity of variance (Bartletts test statistic, P > 0.50). Geographical patterns of genetic diversity were expressed as the mean number of alleles per population and as expected heterozygosity (Neis, 1978) unbiased estimate, He). We tested for signicant variation among regions using one-way ANOVA with all variables meeting the assumptions of normality (Anderson Darling test statistic, P > 0.06) and homogeneity of variance (Bartletts test statistic, P > 0.17). Genetic isolation by distance was tested by comparing the association between matrices of pairwise FST / (1 - FST) and linear geographical distances (km) among populations. The signicance of any patterns were assessed using a MANTEL test (Sokal, 1979) using 10 000 permutations with the software GENETIX, version 4.03 (Belkhir et al., 2001). The shortest geographical distance connecting populations by sea was calculated using MAPSOURCE, version 3.02 (Garmin Corp.).
A) Parvulastra exigua
Among Regions 17%

109

Within Pops 40% Among Pops/ Regions 43%

B) Actinia tenebrosa

Among Regions 12% Among Pops/ Regions 8%

Within Pops 80%

RESULTS
POPULATION
STRUCTURE AND GENETIC SUBDIVISION

C) Meridiastra calcar

Among Regions 0%

Among Pops/ Regions 1%

As expected, our hierarchical AMOVA and F-statistics detected distinct patterns of genetic structuring associated with each life history strategy for the three study species (Fig. 2). For the direct developer, P. exigua, we detected a large and signicant degree of genetic structuring within and between populations and regions (mean FST = 0.60, P < 0.01; Table 1). Up to 43% of the genetic variation could be attributed to differences among populations within each region, whereas 17% of the variation could be attributed to among region differences (Fig. 2A). The remaining variation (40%) was attributed to variation among individuals within each population. The level of genetic differentiation among populations within each of the two regions sampled for P. exigua was similar (FST = 0.475 for the central region and FST = 0.529 for the southern region, respectively; Table 1). For the broadcast spawning/asexually viviparous sea anemone A. tenebrosa, we detected a high degree of genetic differentiation among populations (mean FST = 0.193, P = 0.01; Table 1). However, in comparison to P. exigua, populations of A. tenebrosa within each region showed much lower levels of subdivision, with only 8% of the total variation being attributed to differences among populations within each region. Up to 12% of the variation could be attributed to among

Within Pops 99%

Figure 2. Analysis of molecular variance for three marine species (AC) with contrasting life histories showing the partitioning of genetic variation among regions, among populations within regions, and among individuals within each population.

regions whereas the majority of variation (80%) was attributed to variation among individuals within each population (Fig. 2B). We detected a striking clinal pattern in the frequency of two alleles for the Gpi locus along the northsouth geographic range of A. tenebrosa as previously reported for this species (Ayre et al., 1991). We detected a fast allele that occurred at high frequency, or was completely xed, within populations from northern and central regions, whereas a slow allele occurred at a high frequency in the southern part of the range. Two populations that occurred near the southern centre of its range displayed almost

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

110

C. D. H. SHERMAN ET AL.

Table 1. Hierarchical analysis of standardized genetic variation, showing total variation among all populations (FST), variation among regions (FRT), and total variation among populations within a region (FSR) for three intertidal species along the east coast of Australia FST by region FST Parvulastra exigua Actinia tenebrosa Including all genotypes Including only unique genotypes Meridiastra calcar *P < 0.05, **P < 0.01.
1.0 0.9 0.8 0.7

FRT 0.175** 0.117** 0.114** 0.001

FSR 0.515** 0.086** 0.066** 0.008**

Region 1 0.054* 0.042* 0.002

Region 2 0.475* 0.058 0.074 0.005

Region 3 0.529* 0.117* 0.068* 0.024

0.600** 0.193** 0.173** 0.009**

0.6 0.5 0.4 0.3 0.2 0.1 0.0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19


Fast allele Slow allele

although still signicant, levels of genetic structuring (overall FST = 0.009, P < 0.01; Table 1). We detected no differentiation among regions (0% of the variation) and only 1% of the total variation could be attributed to differences among populations within each region. Almost all (99%) of the variation was attributed to variation among individuals within each population (Fig. 2C), indicating a high degree of connectively between even widely-separated populations. This structure is consistent with high levels of dispersal both within and among regions.

Frequency

Region 1

Region 2

Region 3

PATTERNS
Figure 3. The northsouth clinal variation in the frequency of two alleles along the east coast of Australia for the Gpi locus in the sea anemone Actinia tenebrosa.

OF GEOGRAPHIC VARIATION AND

ISOLATION BY DISTANCE

equal frequencies of each allele (Fig. 3). This distinct clinal variation in the frequency of the two alleles for Gpi1 indicates that this locus (and the closely linked Gpi2) is likely to be under selection. Reanalyses with Gpi1 removed resulted in lower, but still signicant, levels of genetic subdivision among regions and populations (mean FST = 0.122, P < 0.01), although the proportion of variation explained by region dropped by half to 4% of the total, indicating that this locus accounts for a signicant amount of the among region differentiation. For an asexual brooder such as A. tenebrosa, comparisons of levels of genetic diversity or population structure may yield misleading results because of confounding effects of localized clonal recruitment (Ayre, 1983). However, we found that genotypic diversity (the most obvious indicator of clonality) did not vary signicantly among regions when measured either as Ng/N or Go/Ge (ANOVA: F2,16 = 0.17, P > 0.85 and F2,16 = 1.33, P > 0.29, respectively; Table 2). By contrast to P. exigua and A. tenebrosa, the broadcast spawner M. calcar showed much lower,

We detected no consistent pattern of geographic variation in genetic diversity for our three study species. Parvulastra exigua showed no signicant variation among regions in either the mean number of alleles per population (ANOVA: F1,18 = 0.22, P = 0.64) or mean heterozygosity (ANOVA: F1,18 = 0.06, P = 0.82). By contrast, A. tenebrosa and M. calcar both showed signicant variation among regions for the mean number of alleles per population (A. tenebrosa, ANOVA: F2,16 = 6.59, P < 0.01; M. Calcar, ANOVA: F2,23 = 6.00, P < 0.01) but not for expected heterozygosity (A. tenebrosa, ANOVA: F2,16 = 2.35, P = 0.13; M. Calcar, ANOVA: F2,23 = 2.61, P = 0.10). However, for both species, the magnitude of variation was relatively small and, in contrast to our prediction, we did not nd that the central region displayed the lowest allelic diversity. For A. tenebrosa, populations in the southern region showed signicantly greater levels of allelic diversity than the northern region (Turkeys pairwise comparisons, P < 0.02) whereas, for M. calcar, the southern region was signicantly more genetically diverse than the northern and central regions (Turkeys pairwise comparisons, P < 0.02). The relationship between geographical and genetic distance varied among species and was complex. For

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

CONTRASTING PATTERNS OF GENETIC DIFFERENTIATION

111

Table 2. Mean levels of genotypic diversity observed within each of three biogeographic regions for the broadcast spawning/asexual brooding sea anemone Actinia tenebrosa Population Byron Bay Coffs Harbour Fingal Bay Anna Bay Newcastle Cape Banks Bellambi Bass Point Durras Merimula Ben Boyd Green Glades Mallacoota Waratah Bay Cape Paterson Flinders Bicheno Orford Tasman Peninsula N 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 24 Ng 11 23 22 22 17 23 16 18 21 20 21 10 19 15 20 17 21 14 19 Ng/N 0.46 0.96 0.92 0.92 0.71 0.96 0.67 0.75 0.88 0.83 0.88 0.42 0.79 0.63 0.83 0.71 0.88 0.58 0.79 Go 8.2 22.2 20.6 19.2 12.3 22.2 9.9 13.7 18.0 15.6 19.2 5.5 10.7 11.5 15.2 11.5 19.2 5.5 17.0 Ge (SD) 14.5 20.4 22.6 21.9 23.0 23.7 22.6 22.5 22.0 23.1 22.8 22.3 23.7 22.5 22.9 23.1 22.9 22.3 23.2 (3.4) (3.3) (2.3) (2.8) (1.7) (0.8) (2.1) (2.0) (2.6) (1.5) (2.1) (2.3) (2.2) (2.1) (1.8) (1.7) (1.8) (2.4) (1.9) Go/Ge 0.57 1.09 0.91 0.88 0.53 0.93 0.44 0.61 0.82 0.67 0.84 0.25 0.45 0.51 0.66 0.50 0.84 0.25 0.73 P-value < 0.010** NS NS NS < 0.001*** NS < 0.001*** < 0.010** NS < 0.001*** NS < 0.001*** < 0.001*** < 0.001*** < 0.001*** < 0.001*** < 0.020* < 0.001*** < 0.001***

N, number of individuals sampled; Ng, number of unique multi-locus genotypes; Go, observed multi-locus genotypic diversity; Ge, expected multi-locus genotypic diversity for random mating. Signicant values determined after the application of a sequential Bonferroni correction (Rice, 1989). *P < 0.05, **P < 0.01 & ***P < 0.001.

the direct developer P. exigua, we detected a weak but signicant relationship between geographic and genetic distance for the entire set of populations (r2 = 0.08; Mantel test, P = 0.001; Fig. 4A). However, this pattern was not consistent when each region was analysed separately (r2 < 0.05; Mantel test, P > 0.34). For the broadcast spawner/asexual brooder A. tenebrosa, we detected a strong relationship between genetic and geographical distance across all populations (r2 = 0.66; Mantel test, P < 0.001; Fig. 4B). This relationship remained signicant when assessed for populations within the northern region (r2 = 0.59; Mantel test, P = 0.048) but was weaker for populations in the central and southern regions (r2 = 0.03 and r2 = 0.26 respectively; Mantel test, P > 0.08). For the broadcast spawner M. calcar, we detected a weaker but still signicant pattern of isolation by distance among all populations (r2 = 0.09; Mantel test, P = 0.01; Fig. 4C) but, again, this pattern broken down when populations within each region were analysed separately (r2 < 0.01; Mantel test, P > 0.33).

DISCUSSION
The present study supports the prediction that life history is a major determinant of population genetic structure (Wright, 1943; Hedgecock, 1986; Bohonak,

1999; Ayre & Hughes, 2004) even within a complex oceanographic system. We found that population differentiation was inversely related to predicted dispersal ability for the three studied species, and that this showed little variation among three geographic regions, despite great variation in the predicted inuence of ocean currents. This lack of interaction between population structure and the current regime is remarkable because these life histories ranged from that of the direct developing cushion star P. exigua, which only appears likely to be able to disperse widely via rafting (Raff & Byrne, 2006), to the planktonically dispersing cushion star M. calcar, with seasonal episodes of spawning (Byrne, 1992). Despite the prediction that genetic discontinuity will occur coincidently with major biogeographic boundaries (Avise et al., 1987; Avise, 1992; Burton, 1998) due to the disruption of gene ow across such areas, we found no evidence for such a break in any of the three study species at Ninety Mile Beach or across Bass Strait. This contrasts sharply with several earlier studies that have indicated a possible phylogenetic break around Ninety Mile Beach (Ayre et al., 1991; Billingham & Ayre, 1996; Waters & Roy, 2003; Waters, OLoughlin & Roy, 2004; Waters et al., 2005). Although geographic variation in both P. exigua and the asexually vivparous A. tenebrosa do show evi-

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

112

C. D. H. SHERMAN ET AL.

A)
8.0 7.0 6.0

Parvulstra exigua

Fst/(1-Fst)

5.0 4.0 3.0 2.0 1.0 0.0 0 200 400 600 800 1000 1200

Distance (km)

B)
0.6 0.5

Actinia tenebrosa

Fst/(1-Fst)

0.4 0.3 0.2 0.1 0 0 500 1000 1500 2000 2500 3000

Distance (km)

C)
0.10 0.09 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0.00 0 500 1000 1500 2000

Meridiastra calcar

Fst/(1-Fst)

Distance (km)
Figure 4. Patterns of isolation by distance in three intertidal marine invertebrates with contrasting life histories.

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

CONTRASTING PATTERNS OF GENETIC DIFFERENTIATION dence of some genetic differentiation between the central and southern region, closer inspection reveals that this is no greater than that expected given weak evidence of general isolation by distance in P. exigua and a stronger pattern for A. tenebrosa. The low levels of genetic subdivision detected among populations in M. calcar are typical of marine species with high levels of connectivity (Hunt & Ayre, 1989; Todd, Lambert & Thorpe, 1998; Ayre & Hughes, 2000) and are consistent with the lack of isolation by distance revealed by our analysis. The lack of any detectable effect of ocean currents or biogeographic barriers on patterns of genetic subdivision among regions in this species is not surprising considering that the long-lived dispersive larvae should be able to cross potential biogeographic barriers such as Ninety Mile Beach where suitable habitat for settlement is absent. This pattern of genetic homogeneity over large spatial scales is consistent with the genetic structure reported for other asteroids with dispersive feeding larvae (Williams & Benzie, 1996; Benzie & Wakeford, 1997; Colgan et al., 2005). By contrast, P. exigua populations are highly genetically subdivided at all spatial scales and populations appear to be effectively closed and self-seeding. This supports the study by Hunt (1993) reporting high levels of genetic differentiation (FST = 0.46) over distances of 230 km in this species. Similarly, mitochondrial DNA analysis by Waters & Roy (2004) and Colgan et al. (2005) found little evidence of connectivity between NSW, South Australian, and Tasmania regions. Thus, a distance of several kilometres appears to be just as effective a barrier to dispersal as larger areas of unsuitable habitat such as that seen at Ninety Mile Beach. Nevertheless, P. exigua has a very extensive geographical range that spans areas such as Ninety Mile Beach, which indicates that larger-scale dispersal through rafting of larvae and/or egg masses must occur at least occasionally. However, we did not detect the decrease in diversity for populations on either side of the biogeographic barrier at Ninety Mile Beach that would be predicted if populations had been established by rare founder events. This suggests that much of the underlying population structure, and even species distribution, may reect historical patterns of gene ow and an earlier range expansion in this species (Waters & Roy, 2004; Colgan et al., 2005). Thus, current patterns of genetic differentiation and diversity may not reect contemporary barriers to gene ow. Further genetic analysis using a range of genetic markers and more in-depth sampling of populations within each region may prove useful in discerning the effects of historical versus contemporary gene ow in this species (Hedgecock, 1986; Avise et al., 1987; Grosberg & Cunningham, 2001).

113

We detected relatively high levels of genetic differentiation among populations of A. tenebrosa along the east coast of mainland Australia and Tasmania, which displayed a distinct pattern of isolation by distance. Such increasing genetic differentiation with geographical distance is likely to reect the effects of dispersal of propagules among only neighbouring populations (Wright, 1943, 1969; Slatkin, 1987, 1993). However, our data do not support previous reports of a genetic discontinuity between NSW and Victoria concordant with the proposed biogeographic barrier focused around Ninety Mile Beach. Similarly, our data do not support our prediction that the Bass Strait will act as a signicant barrier to dispersal among Australian mainland and Tasmanian populations. The level of genetic differentiation among populations spanning the NSW and Victorian regions was no greater than that detected among other geographic regions. Previous evidence for a discontinuity around Ninety Mile Beach in this and another anemone species is likely to reect the inuence of a temperature dependent Gpi polymorphism in those studies (Ayre et al., 1991; Billingham & Ayre, 1996). The Gpi locus in A. tenebrosa shows steep clinal variation with the fast allele found at high frequencies or completely xed in northern and central NSW populations, whereas the slow allele occurs at a much higher frequency in southern Victorian and Tasmanian populations (Fig. 3). Similar reports of a latitudinal cline in the frequency of some allozyme loci have been reported in a number of marine invertebrates (Shick, Hoffmann & Lamb, 1979; Koehn, Newell & Immermann, 1980; Hoffmann, 1981; Hedgecock, 1986). Reanalysis of levels of genetic differentiation with Gpi removed still however revealed signicant levels of genetic subdivision among populations and regions. The use of different loci and types of markers (i.e. allozyme and microsatellites for A. tenebrosa) may potentially bias results due to differences in apparent variation, with more variable microsatellite markers potentially showing higher levels of genetic structure than allozymes. Despite this potential bias, we found that populations of A. tenebrosa did not show higher levels of genetic structure than the direct developing cushion start P. exigua, suggesting that the higher levels of variability of the microsatellite markers has not overwhelmed differences due to dispersal potential. The apparent lack of genetic discontinuity at a potentially major biogeographical barrier indicates that, for nuclear loci, neither the disruption of gene ow by oceanographic currents, nor the availability of suitable substrate are sufficient to result in genetic differentiation among populations spanning this region for some species with planktonically dispersed larvae. This is consistent with a growing number of

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

114

C. D. H. SHERMAN ET AL.
variation in clonal structure in a reef-building Caribbean coral, Acropora palmata. Ecological Monographs 76: 503 519. Belkhir K, Borsa P, Chikhi L, Raufaste N, Bonhomme F. 2001. GENETIX 4.02, logiciel sous Windows TM pour la gntique des populations. Montpellier: Laboratoire Gnome, Populations, Interactions, CNRS UMR 5000, Universit de Montpellier II. Bennett I, Pope EC. 1953. Intertidal zonation of the exposed rocky shores of Victoria, together with a rearrangement of the biogeographical provences of temperate Australasian shores. Australian Journal of Marine & Freshwater Research 4: 105159. Benzie JAH, Wakeford M. 1997. Genetic structure of crownof-thorns starsh (Acanthaster planci) on the Great Barrier Reef, Australia: comparison of two sets of outbreak populations occurring ten years apart. Marine Biology 129: 149 157. Benzie JAH, Williams ST. 1995. Gene Flow among giant clam (Tridacna gigas) populations in Pacic does not parallel ocean circulation. Marine Biology 123: 781787. Billingham M, Ayre DJ. 1996. Genetic subdivision in the subtidal, clonal sea anemone Anthothoe albocincta. Marine Biology 125: 153163. Billingham MR, Reusch TBH, Alberto F, Serrao EA. 2003. Is asexual reproduction more important at geographical limits? A genetic study of the seagrass Zostera marina in the Ria Formosa, Portugal. Marine Ecology Progress Series 265: 7783. Bohonak AJ. 1999. Dispersal, gene ow, and population structure. Quarterly Review of Biology 74: 2145. Burton RS. 1998. Intraspecic phylogeography across the Point Conception biogeographic boundary. Evolution 52: 734745. Byrne M. 1992. Reproduction of sympatric populations of Patiriella gunnii, P. calcar and P. exigua in New South Wales, Asterinid seastars with direct development. Marine Biology 114: 297316. Chiswell SM, Toole J, Church J. 1997. Transports across the Tasman Sea from WOCE repeat sections: the East Australian Current 1990-94. New Zealand Journal of Marine and Freshwater Research 31: 469475. Colgan DJ, Byrne M, Rickard E, Castro LR. 2005. Limited nucleotide divergence over large spatial scales in the Asterinid sea star Patiriella exigua. Marine Biology 146: 263270. Edmands S, Moberg PE, Burton RS. 1996. Allozyme and mitochondrial DNA evidence of population subdivision in the purple sea urchin Strongylocentrotus purpuratus. Marine Biology 126: 443450. Excoffier L, Laval G, Schneider S. 2005. Arlequin ver. 3.0: an integrated software package for population genetics data analysis. Evolutionary Bioinformatics Online 1: 4750. Excoffier L, Smouse PE, Quattro JM. 1992. Analysis of molecular variance inferred from metric distances among DNA haplotypes application to human mitochondrial-DNA restriction data. Genetics 131: 479491. Farquhar H. 1898. Preliminary account of some New

studies that have shown that major current patterns are not always concordant with patterns of genetic differentiation (Benzie & Williams, 1995; Shulman & Bermingham, 1995; Edmands, Moberg & Burton, 1996; Hellberg, 1996; Burton, 1998). Our data support the widely held view that marine species with widely dispersed larvae show little genetic structuring over large distances whereas those with limited dispersal show genetic structuring over much ner spatial scales (Hedgecock, 1986; Bohonak, 1999; Ayre & Hughes, 2000; Sherman, 2006). Our data suggest that life history is the major determinant of genetic structure of populations and that the complexity of the oceanographic current and potential biogeographical barriers along the east coast of Australia do not appear to cause any obvious disruption of genetic connectivity. Clearly, comparable data for a phylogentically diverse set of species are needed to test the generality of our conclusions.

ACKNOWLEDGEMENTS
This work was supported by the University of Wollongong through a Postgraduate Scholarship to C.D.H.S. and A.H., the Institute for Conservation Biology and an ARC Discovery Grant to D.J.A.

REFERENCES
Avise JC. 1992. Molecular population structure and the biogeographic history of a regional fauna a case history with lessons for conservation biology. Oikos 63: 6276. Avise JC, Arnold J, Ball RM, Bermingham E, Lamb T, Neigel JE, Reeb CA, Saunders NC. 1987. Intraspecic phylogeography the mitochondrial DNA bridge between population genetics and systematics. Annual Review of Ecology and Systematics 18: 489522. Ayre DJ. 1982. Inter genotype aggression in the solitary sea anemone Actinia tenebrosa. Marine Biology 68: 199206. Ayre DJ. 1983. The distribution of the sea anemone Actinia tenebrosa in southwestern Australia. Western Australian Naturalist 15: 136140. Ayre DJ. 1984. The effects of sexual and asexual reproduction on geographic variation in the sea anemone Actinia tenebrosa. Oecologia 62: 222229. Ayre DJ, Hughes TP. 2000. Genotypic diversity and gene ow in brooding and spawning corals along the Great Barrier Reef, Australia. Evolution 54: 15901605. Ayre DJ, Hughes TP. 2004. Climate change, genotypic diversity and gene ow in reef-building corals. Ecology Letters 7: 273278. Ayre DJ, Read J, Wishart J. 1991. Genetic subdivision within the eastern Australian population of the sea anemone Actinia tenebrosa. Marine Biology 109: 379390. Baums IB, Miller MW, Hellberg ME. 2006. Geographic

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

CONTRASTING PATTERNS OF GENETIC DIFFERENTIATION


Zealand Actiniaria. Journal of the Linnean Society of London (Zoology) 26: 527536. Grosberg RK, Cunningham CW. 2001. Genetic structure in the sea: from populations to communities. In: Bertness MD, Gaines SD, Hay ME, eds. Marine community ecology. Sunderland, MA: Sinauer Associates, 6184. Hedgecock D. 1986. Is gene ow from pelagic larval dispersal important in the adaption and evolution of marine invertebrates? Bulletin of Marine Science 39: 550564. Hellberg ME. 1996. Dependence of gene ow on geographic distance in two solitary corals with different larval dispersal capabilities. Evolution 50: 11671175. Hidas EZ, Costa TL, Ayre DJ, Minchinton TE. 2007. Is the species composition of rocky intertidal invertebrates across a biogeographic barrier in southeast Australia related to their potential for dispersal? Marine and Freshwater Research 58: 835842. Hoffmann RJ. 1981. The evolutionary genetics of Metridium senile. I. Kinetic differences in the phosphoglucose isomerase allozymes. Biochemical Genetics 19: 129144. Hunt A. 1993. Effects of contrasting dispersal strategies on populations of two species of intertidal starsh, Patiriella calcar and Patiriella exigua. PhD Thesis, School of Biological Sciences: University of Wollongong. Hunt A, Ayre DJ. 1989. Population structure in the sexually reproducing sea anemone Oulactis muscosa. Marine Biology 102: 537544. Koehn RK, Newell RIE, Immermann F. 1980. Maintenance of an aminopeptidase allele frequency cline by natural-selection. Proceedings of the National Academy of Sciences of the United States of America 77: 53855389. Lamarck J. 1816. Histoire Naturelle des Animaux Sans Vertbres. Paris: Verdire. Lawson-Kerr C, Anderson DT. 1978. Reproduction, spawning and development of the starsh Patiriella exigua (Lamarck) (Asteroidea, Asterinidae) and some comparisons with P. calcar (Lamarck). Australian Journal of Marine & Freshwater Research 29: 4553. Lessios HA, Kessing BD, Pearse JS. 2001. Population structure and speciation in tropical seas: Global phylogeography of the sea urchin Diadema. Evolution 55: 955 975. Murray-Jones SE, Ayre DJ. 1997. High levels of gene ow in the surf bivalve Donax deltoides (Bivalvia: Donacidae) on the east coast of Australia. Marine Biology 128: 8389. Nei M. 1978. Estimation of average heterozygosity and genetic distance from a small number of individuals. Genetics 89: 583590. Palumbi SR. 1994. Genetic divergence, reproductive isolation, and marine speciation. Annual Review of Ecology and Systematics 25: 547572. Raff RA, Byrne M. 2006. The active evolutionary lives of echinoderm larvae. Heredity 97: 244252. Raymond M, Rousset F. 1995. Genepop (version 1.2): population genetics software for exact tests and ecumenicism. Journal of Heredity 86: 248249. Rice WR. 1989. Analyzing tables of statistical tests. Evolution 43: 223225.

115

Ridgway KR, Godfrey JS. 1997. Seasonal cycle of the East Australian Current. Journal of Geophysical ResearchOceans 102: 2292122936. Sherman CDH. 2006. The importance of ne-scale environmental heterogeneity in determining levels of genotypic diversity and local adaptation. PhD Thesis, School of Biological Sciences, Wollongong, Australia. Sherman CDH, Peucker AJ, Ayre DJ. 2007. Do reproductive tactics vary with habitat heterogeneity in the intertidal sea anemone Actinia tenebrosa? Journal of Experimental Marine Biology and Ecology 340: 259267. Shick JM, Hoffmann RJ, Lamb AN. 1979. Asexual reproduction, population structure, and genotype environment interactions in sea anemones. American Zoologist 19: 699 713. Shulman MJ, Bermingham E. 1995. Early life histories, ocean currents, and the population genetics of Caribbean reef shes. Evolution 49: 897910. Slatkin M. 1987. Gene ow and geographic structure of natural populations. Science 236: 787792. Slatkin M. 1993. Isolation by distance in equilibrium and nonequilibrium populations. Evolution 47: 264279. Sokal RR. 1979. Testing statistical signicance of geographic variation patterns. Systematic Zoology 28: 227232. Stoddart JA. 1988. Coral populations fringing islands larval connections. Australian Journal of Marine and Freshwater Research 39: 109116. Stoddart JA, Taylor JF. 1988. Genotypic diversity estimation and prediction in samples. Genetics 118: 705711. Strathmann RR. 1985. Feeding and nonfeeding larval development and life-history evolution in marine invertebrates. Annual Review of Ecology and Systematics 16: 339361. Todd CD, Lambert WJ, Thorpe JP. 1998. The genetic structure of intertidal populations of two species of nudibranch molluscs with planktotrophic and pelagic lecithotrophic larval stages: are pelagic larvae for dispersal? Journal of Experimental Marine Biology and Ecology 228: 128. Waters JM, King TM, OLoughlin PM, Spencer HG. 2005. Phylogeographical disjunction in abundant highdispersal littoral gastropods. Molecular Ecology 14: 2789 2802. Waters JM, OLoughlin PM, Roy MS. 2004. Cladogenesis in a starsh species complex from southern Australia: evidence for vicariant speciation? Molecular Phylogenetics and Evolution 32: 236245. Waters JM, Roy MS. 2003. Marine biogeography of southern Australia: phylogeographical structure in a temperate seastar. Journal of Biogeography 30: 17871796. Waters JM, Roy MS. 2004. Out of Africa: the slow train to Australasia. Systematic Biology 53: 1824. Weir BS. 1979. Inferences about linkage disequilibrium. Biometrics 35: 235254. Weir BS, Cockerham CC. 1984. Estimating F-statistics for the analysis of population structure. Evolution 38: 1358 1370. Williams ST, Benzie JAH. 1993. Genetic consequences of

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

116

C. D. H. SHERMAN ET AL.
Oceans, revealed by allozyme electrophoresis. Marine Biology 126: 99107. Wright S. 1943. Isolation by distance. Genetics 28: 139156. Wright S. 1969. The evolution and genetics of populations. Vol. 2. The theory of gene frequencies. Chicago, IL: University of Chicago Press.

long larval life in the starsh Linckia laevigata (Echinodermata, Asteroidea) on the Great Barrier Reef. Marine Biology 117: 7177. Williams ST, Benzie JAH. 1996. Genetic uniformity of widely separated populations of the coral reef starsh Linckia laevigata from the East Indian and West Pacic

2008 The Linnean Society of London, Biological Journal of the Linnean Society, 2008, 95, 106116

You might also like