23
Chapter 10 Gene flow and the evolutionary ecology of passively dispersing aquatic invertebrates Beth Okamura and Joanna R. Freeland Introduction Apart from a relatively small number of ancient inland bodies of water, most fresh- water lakes originated during the last glaciation and will eventually disappear due to processes such as infilling with sediments and encroachment of marginal vegetation (Wetzel 1975). Therefore, to ensure long-term persistence, most freshwater organ- isms must at least occasionally disperse to and colonize new sites, and the wide- spread distribution of many species is indicative of such events (Darwin 1859; Pennak 1989). However, as dispersal is notoriously difficult to assess, its extent and frequency are poorly understood. This ignorance significantly compromises our understanding of both population structure and population dynamics. This chapter will review the evidence for, and assess the biological significance of, intersite dispersal and gene flow by invertebrates that inhabit lakes and ponds. We begin by broadly considering how variation in the frequency and scale of dispersal affects population and community structures. We briefly review dispersal mecha- nisms of freshwater invertebrates, and then focus on our own research programme on dispersal, gene flow and the metapopulation ecology of freshwater bryozoans. This is followed by a review of dispersal in freshwater zooplankton,an approach that allows us to compare and contrast the significance and extent of dispersal and gene flow for groups that differ in ecology and life history. Frequency and scale of dispersal Research over the last decade on the spatial ecology of populations and communities has revealed that a profound influence may be exerted by the frequency and scale of dispersal. Much of the research at the population level has focused on metapopula- tion dynamics, whereas community ecologists have focused on the relative impor- tance of regional verus local influences on patterns of species richness. Dispersal and metapopulation structure Since many organisms occur in a series of discrete and isolated sites, the metapopu- 194

Chapter 10 Gene flow and the evolutionary ecology of passively

  • Upload
    dinhque

  • View
    218

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Chapter 10 Gene flow and the evolutionary ecology of passively

Chapter 10Gene flow and the evolutionary ecology ofpassively dispersing aquatic invertebrates

Beth Okamura and Joanna R. Freeland

IntroductionApart from a relatively small number of ancient inland bodies of water, most fresh-water lakes originated during the last glaciation and will eventually disappear due toprocesses such as infilling with sediments and encroachment of marginal vegetation(Wetzel 1975). Therefore, to ensure long-term persistence, most freshwater organ-isms must at least occasionally disperse to and colonize new sites, and the wide-spread distribution of many species is indicative of such events (Darwin 1859;Pennak 1989). However, as dispersal is notoriously difficult to assess, its extent andfrequency are poorly understood. This ignorance significantly compromises ourunderstanding of both population structure and population dynamics.

This chapter will review the evidence for, and assess the biological significance of,intersite dispersal and gene flow by invertebrates that inhabit lakes and ponds. Webegin by broadly considering how variation in the frequency and scale of dispersalaffects population and community structures. We briefly review dispersal mecha-nisms of freshwater invertebrates, and then focus on our own research programmeon dispersal, gene flow and the metapopulation ecology of freshwater bryozoans.This is followed by a review of dispersal in freshwater zooplankton,an approach thatallows us to compare and contrast the significance and extent of dispersal and geneflow for groups that differ in ecology and life history.

Frequency and scale of dispersalResearch over the last decade on the spatial ecology of populations and communitieshas revealed that a profound influence may be exerted by the frequency and scale ofdispersal. Much of the research at the population level has focused on metapopula-tion dynamics, whereas community ecologists have focused on the relative impor-tance of regional verus local influences on patterns of species richness.

Dispersal and metapopulation structureSince many organisms occur in a series of discrete and isolated sites, the metapopu-

194

Page 2: Chapter 10 Gene flow and the evolutionary ecology of passively

lation concept (Levins 1969, 1970) has been increasingly applied by ecologists to describe the dynamics of subdivided populations. Levins’ classic metapopulationmodel demonstrated that, in a collection of subdivided populations, the fraction ofhabitat patches occupied at any one time results from a balance of the rate at whichlocal populations go extinct and the rate of colonization of empty patches (Fig. 10.1)(Hanski 1989). Since all local populations are subject to extinction, persistence ap-plies only at regional scales (Hanski & Simberloff 1997). Dispersal lies at the heart ofmetapopulation dynamics since it determines which habitat patches collectivelycomprise a metapopulation at any given time.

Although metapopulation dynamics are believed to apply to many populations,conformation to the classic metapopulation model appears to be relatively rare(Harrison & Taylor 1997; Bullock et al., this volume). Further refinements of theclassic model include effects such as variation in patch size or quality on extinctionprobabilities (e.g. island–mainland and source–sink metapopulations), the influ-ence of interspecific interactions on rates of local extinction and colonization, andspatially dependent variation in the probability of dispersal between sites (e.g. step-

195

Occupied sites

Unoccupied sites

Figure 10.1 Schematic representation of a classic metapopulation. Populations within localsites go extinct and are recolonized through dispersal (shown by arrows) from othercurrently occupied sites within the metapopulation. The pattern of site occupation variesover time with dispersal playing a crucial role in linking subpopulations within themetapopulation.

Page 3: Chapter 10 Gene flow and the evolutionary ecology of passively

ping stone and other spatially explicit models) (for reviews see Hanski & Simberloff1997; Hanski 1999). The latter developments can account for effects of the vagility ofspecies and landscape structure. Thus, whether a series of patches represent sepa-rate, independent populations, true metapopulations or ‘patchy populations’ (uni-tary populations within which there are discrete patches of resources; Menéndez &Thomas 2000) will be influenced by the frequency and scale of dispersal across thelandscape.

The frequency and scale of dispersal and accompanying gene flow will also influ-ence patterns of genetic differentiation (Slatkin 1987). Thus, patchy populationsshould converge towards panmixia given frequent gene flow, and there will be littlegenetic differentiation among local populations. At intermediate levels, gene flowwill link subpopulations within a metapopulation, and moderate levels of geneticdifferentiation are expected due to, for example, founder effects, genetic drift andlocal selection. Finally, in independent populations genetic differentiation shouldultimately lead to speciation.

Dispersal and patterns of species richnessConsideration of dispersal at the community level provides the basis for regionalversus local patterns of species richness. Patterns of species composition among sitesmay reflect the extent to which dispersal and local interactions limit local diversity(for reviews see Cornell & Karlson 1997; Srivastava 1999). If dispersal rates greatlyexceed local extinction probabilities for many species, then nearly all species in theregion that are capable of invading the site should be present. Local processes willthen determine species diversity and composition within patches. In this case, localspecies richness will approach an upper asymptote with increasing regional richness(Fig. 10.2) (Cornell & Lawton 1992). However, if dispersal is rare, then many speciesmay be absent. Local species composition in this case will depend on a site’s historyof colonization,and local interactions may be relatively unimportant (Ricklefs 1987;Cornell & Lawton 1992; Cornell & Karlson 1997). Species composition will then beundersaturated, and local species richness will increase continuously with increas-ing richness of the regional pool of species (Fig. 10.2). Such communities are con-sidered to be dispersal limited and under strong regional control. Linear, positiverelationships in support of dispersal limitation have been discerned in a diversity ofsystems in marine, freshwater and terrestrial environments (Cornell & Karlson1997; Srivastava 1999).

How do freshwater invertebrates disperse?Intersite dispersal of freshwater invertebrates can be achieved by active or passivemeans (for review see Bilton et al. 2001). Active dispersal occurs through self-generated movements of organisms, mainly exemplified by aerial flight of adult in-sects. Passive dispersal is achieved by means of external agents, including floods,wind and animal vectors. Most non-insect invertebrates of lakes and ponds colonizenew sites through the passive dispersal of small, dormant propagules (Williams

196 . . .

Page 4: Chapter 10 Gene flow and the evolutionary ecology of passively

1987). These propagules are typically highly resistant to desiccation and extremes oftemperature and often show apparent adaptations to increase the likelihood of at-tachment to animal vectors. Thus, many propagules display features such as hooks,spines and sticky surfaces (Fig. 10.3); avoidance of deposition in bottom sedimentsthrough buoyancy; and accumulations of large numbers at times that coincide withpeaks of waterfowl annual migrations (Bilton et al. 2001).

Dispersal, life history and population genetic structure in freshwater bryozoans

The life history of freshwater bryozoansFreshwater bryozoans are benthic, hermaphroditic, colonial invertebrates and arecommon, but overlooked, residents of freshwater habitats (Wood 1991). Their lifecycle involves a high degree of clonal reproduction through colony growth by bud-ding of new zooids, colony fission or fragmentation, and the asexual production oflarge numbers of resistant propagules, called statoblasts. Statoblasts represent bothan overwintering and dispersal stage (Fig. 10.3). Small colonies emerge from stato-blasts when favourable conditions return in the following growing season. Sexual re-production is brief in duration and results in short-lived, swimming larvae that can

197

Unsaturatedcommunities

Saturatedcommunities

Regional species richness

Loca

l spe

cies

ric

hnes

s

Figure 10.2 In unsaturated communities local species richness is predicted to be a fixedproportion of regional species richness. In saturated communities, local richness mayincrease with regional richness but will reach an upper limit at higher regional richness.(Adapted from Srivastava 1999.)

Page 5: Chapter 10 Gene flow and the evolutionary ecology of passively

disperse within a site before metamorphosing into a small colony. The early sea-sonal occurrence of sex in bryozoans is unusual. The majority of freshwater inverte-brate groups undergo a late season sexual phase which is viewed to be adaptive inthat it results in the production of maximal genetic variation at a time that coincideswith unpredictable conditions (Lynch & Spitze 1994). Thus, when conditions deteriorate, zooplankton taxa undergo a sexual phase to produce dormant, resistantembryos that will hatch out the following season.

Population genetic studies of Cristatella mucedoThe bryozoan Cristatella mucedo occurs as elongate,gelatinous colonies in lakes andponds throughout the Holarctic. This widespread distribution has allowed our population genetics studies to be conducted over large spatial scales; thus approxi-mately 30 colonies of C. mucedo were collected from each of 12 water bodies across atransect of approximately 2500 km in northwest Europe and from each of eightwater bodies across a transect of approximately 1500 km in central North America(Fig. 10.4). Each colony was collected from separate substrata to avoid collectingreplicate genotypes from aggregations of genetically identical colonies that can develop through colonization by a single statoblast or larva and subsequent colonygrowth and fission. Note,however, that dissemination of floating statoblasts can po-tentially spread the same clonal genotype throughout a site. Microsatellite profiles

198 . . .

Figure 10.3 Scanning electron micrograph of statoblasts of Cristatella mucedo. Scale bar = 400 mm.

Page 6: Chapter 10 Gene flow and the evolutionary ecology of passively

199

Figure 10.4 (a) Map showing the location of 12 sites sampled for Cristatella mucedo innorthwest Europe with track of main waterfowl annual migratory route traversing theregion sampled shown by arrow. (b) Map showing the location of eight sites sampled incentral North America with pathways of numerous, divergent annual migratory waterfowlroutes traversing the region shown by arrows.

Page 7: Chapter 10 Gene flow and the evolutionary ecology of passively

were generated for all colonies, and a subset of these was further genetically charac-terized using mitochondrial DNA sequence data (for further details, see Freeland et al. 2000c).

Analysis of gene flow and population structure using microsatellitesWithin each continental region, we characterized populations based on five micro-satellite loci (Freeland et al. 2000a, 2000b). However, only three of these loci werereadily amplified from both North American and European C. mucedo (Freeland et al. 2000c), and we base much of our discussion on patterns revealed by these threeloci. This selective presentation of data is justified because direct comparisons ofmicrosatellite data from different loci may be hampered by inherently variable mu-tation rates. Furthermore, the results of separate population genetic analyses basedon the full set of five microsatellite loci (Freeland et al. 2000a, 2000b) provided simi-lar conclusions to analyses based on the three common loci (Freeland et al. 2000c).To support this latter statement, we refer to conclusions based on the full data sets(Freeland et al. 2000a, 2000b) when appropriate.

Traditional means of gaining evidence for dispersal have relied on estimates ofgene flow (Nm) based on FST approaches. However, it is now widely appreciated thatsuch estimates may be flawed owing to violation of assumptions such as populationequilibrium, no spatial variation in migration rates and population sizes, and nomutation (Whitlock & McCauley 1999; Raybould et al., this volume). In view ofthese problems we conducted non-parametric discriminant function analyses as anindependent means of obtaining evidence for gene flow amongst populations within each continent (Freeland et al. 2000c). These analyses classified 32.5% ofEuropean colonies (representing 51.3% of multilocus genotypes, MLGs) to popu-lations other than the one from which they were collected, compared to 18.3% ofcolonies (representing 24% of MLGs) in North America. When discriminant func-tion analyses were performed on the full data sets of five microsatellite loci, 14% of European colonies were misclassified and only 8% were misclassified in NorthAmerica (Freeland et al. 2000a, 2000b).

The occurrence of identical MLGs (identified by the three microsatellites) in mul-tiple populations provided further support for higher levels of gene flow in Europecompared to North America. In Europe,nine MLGs were each found in two popula-tions and one MLG was found in four populations. In North America, only twoMLGs were found in multiple (in both cases two) populations. We should point outthe possibility that these MLG distributions would be uninformative about levels ofgene flow if the MLGs arose by processes other than recent common ancestry. How-ever, in an earlier study, we specifically tested this possibility by conducting a com-puter simulation to assess the likelihood that two MLGs (based on five microsatelliteloci), found in European lakes separated by 700 km,resulted from a chance recombi-nation of alleles (Freeland et al. 2000a). The likelihood of such a chance recombina-tion of alleles at the two sites was exceedingly small (P < 0.001). While we did notconduct such simulations to determine the likelihoods that all the MLGs arose bychance in the various sites discussed here, our previous simulation study and con-

200 . . .

Page 8: Chapter 10 Gene flow and the evolutionary ecology of passively

cordance with other inferences of gene flow (see below) together provide evidencethat MLG distributions give a useful indication of relative levels of gene flowamongst the populations sampled.

Although the potential inaccuracies of FST-based estimates of gene flow must beborn in mind,Nm values conformed with the above conclusions of greater gene flowamongst European than North American populations. Overall levels of Nm basedon r (an analogue of FST developed for use with microsatellite data; Goodman 1997)were higher among European (range = 0.11 - 14.18, mean ± SD = 1.21 ± 0.31) thanNorth American populations (range = 0.03 - 0.87, mean ± SD = 0.20 ± 0.04). WhenNm was derived from FST, the difference was reduced but values were still con-siderably higher for European populations (Europe: range = 0.01 - 4.21, mean ± SD= 0.70 ±0.007;North America: range =0.19 - 0.73,mean ±SD =0.44 ±0.03). The rel-atively higher Nm estimates for European populations are mirrored by analysesbased on the five microsatellite loci for each continent (Freeland et al. 2000a,2000b).

In summary, consistent results provided by our three separate approaches (discriminant function assignments, distributions of MLGs and Nm estimates) together provide strong evidence for ongoing gene flow amongst populations innorthwest Europe and considerably reduced to non-existent levels of gene flowamongst populations in central North America.

As estimates of Nm are inversely correlated with values of FST or r, the higher estimates of gene flow in Europe are naturally equated with lower, but none the less substantial, levels of population subdivision in Europe compared with NorthAmerica. High levels of genetic differentiation for populations in both continentswere similarly supported by analysis of the full data set of five microsatellite loci,withoverall values being lower in Europe (Freeland 2000a, 2000b). Mantel tests showedno pattern of isolation by distance in North America or Europe when either r or FST

values were used (Fig. 10.5). The lack of isolation by distance was also supported byanalyses of full data sets based on five microsatellite loci (Freeland et al. 2000a,2000b).

Data sets based on both three and five microsatellite loci revealed that none of the39 populations on either continent were in Hardy–Weinberg equilibrium (Freelandet al. 2000a, 2000b, 2000c). If random sampling from clonally reproducing popula-tions was responsible for an absence of Hardy–Weinberg equilibrium then we wouldexpect to see instances of both observed heterozygosity (Ho) deficits and excesses.However, in all 39 C. mucedo populations, deviations from Hardy–Weinberg pro-portions were consistently due to Ho deficits, which were reflected in high estimatesof FIS (Freeland et al. 2000a, 2000b, 2000c). Both inbreeding and the Wahlund effect(see later) are likely to contribute to low Ho values. Self-fertilization and inbreedingmay be promoted by the short distances traversed by sperm spawned in the watercolumn,the development of clonal aggregations through colony growth and fission,and the sessile nature of colonies. However,microsatellite analysis has indicated thatover half of the larval progeny produced by representative colonies from one site arethe products of cross-fertilization since at least one allele was not present in parentalcolonies (J. Freeland, unpubl. data).

201

Page 9: Chapter 10 Gene flow and the evolutionary ecology of passively

Microsatellite analysis revealed similar levels of within-population diversity inEurope and North America as measured by values of He (expected heterozygosity)and I (Shannon–Weaver diversity index) (Freeland et al. 2000a, 2000b, 2000c).There was a relatively higher clonal diversity (number of unique MLGs/number ofcolonies) in North America as a whole compared to Europe (c2 = 13.281, df = 1, P <0.001) as well as a higher allelic diversity (the number of alleles/number of coloniessampled) (z = 4.239; P £ 0.001) (Fig. 10.6).

Analysis of historical dispersal events through mtDNA sequence dataWe complemented our microsatellite analyses by sequencing a region of 16S rDNA(479–480 base pairs) from three randomly chosen MLGs per site, identified from thefull set of five microsatellites.We also sequenced single colonies from 10 other NorthAmerican sites and from five other European sites from which only a small numberof colonies were collected.

The 34 North American colonies yielded 16 haplotypes with a maximum sequence divergence of 0.036 (Freeland et al. 2000c). The 41 European coloniesyielded only three haplotypes with a maximum sequence divergence of 0.006 (Freeland et al. 2000c) (Fig. 10.7). The maximum divergence between North American and European sequences was 0.026. The relative haplotype diversity persite was significantly greater in North America than in Europe (t = 5.01, df = 18, P <0.001) (Fig. 10.8a), but the mean number of sites per haplotype was significantlygreater in Europe (analysis of log-transformed data: t = 2.93, df = 17, P = 0.009) (Fig.10.8b).

202 . . .

Europe

N. America

8

6

4

2

0

0 2 4 6 8 10

ln distance

r/(1

– r

)

Figure 10.5 Genetic relatedness (as r/(1 - r) values) versus distances (as ln distance values)between sites for European and North American populations of Cristatella mucedo.

Page 10: Chapter 10 Gene flow and the evolutionary ecology of passively

203

N. America

Europe

Clonal diversity Relative allelicdiversity

0.6

0.4

0.2

0

Div

ersi

ty

Figure 10.6 Clonal diversity (unique multilocus genotypes/number of colonies) and relativeallelic diversity (number of alleles/number of colonies sampled) in North American andEuropean populations of Cristatella mucedo.

H5

H8H4

H6

H7

H11

H19H18

H17H12

H16

H15H13

H14

H9

H3

H1

H10

H2

93

85

89

74

65

6995

European

0.01

Figure 10.7 Scaled maximum likelihood tree of the 19 mitochondrial haplotypes (H1–H19)of Cristatella mucedo populations sampled in Europe and North America. Bootstrap values>50 are provided above branches. Note that the three haplotypes from Europe (H17–H19)cluster within the range of divergences of North American haplotypes (H1–H16).

Page 11: Chapter 10 Gene flow and the evolutionary ecology of passively

204 . . .

1.0

0.8

0.6

0.4

0.2

0.0Europe N. America

Rel

ativ

e ha

plot

ype

dive

rsity

/site

(a)

1.2

0.8

0.4

0.0N. America EuropeM

ean

no. s

ites/

hapl

otyp

e

(b)

Figure 10.8 (a) Relative haplotype diversity for North American and European populations.Data are plotted as the relative diversity of haplotypes in sites for which haplotypes wereobtained for three multilocus genotypes (see text for further discussion). (b) Mean numberof sites per haplotype for North American and European populations of Cristatella mucedo.Data plotted as log (x + 1).

The degree of divergence amongst the haplotype lineages in North America indicates that many of these haplotypes must have been maintained throughoutglacial–interglacial cycles, as much of the region sampled was covered by ice 10 000–20 000 years ago. The co-occurrence of divergent mitochondrial lineageswithin the same population today must reflect repeated postglacial colonizationevents from glacial refugia, occasional dispersal during postglacial range extensionor ongoing dispersal that was not detected by our genetic analyses (see discussionbelow).

The widespread distribution of haplotypes in Europe may be interpreted as affir-mation that levels of gene flow are higher on that continent. However, as the 16SrDNA sequence is expected to evolve at a substantially slower rate than microsatelliteloci, biogeographical patterns inferred from mitochondrial sequences may reflecthistorical events. The paucity of mitochondrial haplotypes in Europe comparedwith North America and their relatively low degree of divergence strongly suggest ahistorical bottleneck, possibly following intercontinental dispersal from NorthAmerica. This direction of dispersal is supported by the greater similarity of someNorth American haplotypes to European haplotypes than to other North Americansequences (Fig. 10.7). Intercontinental dispersal events have also been identified in zooplankton taxa (Berg & Garton 1994; Taylor et al. 1996;Weider et al. 1999a). Analternative scenario, that relatively few haplotypes were maintained in Europethroughout the glacial–interglacial cycles, is unlikely as this would be reflected by a deep, rather than shallow, coalescence of the European and North American haplotypes.

A recent common ancestor shared amongst European haplotypes could explainthe genetic similarity among sites, thereby confounding estimates of gene flow.However, common ancestry is unlikely to unduly influence interpretations of ongo-ing population connectivity in northwestern Europe because the microsatellite datareveal genetically distinct populations with levels of genetic diversity similar to those

Page 12: Chapter 10 Gene flow and the evolutionary ecology of passively

in North America,as well as a number of unique alleles. While the microsatellite locimay be expected to retain higher levels of diversity following a population bottle-neck as compared with mitochondrial DNA, as nuclear DNA is effectively four times the size of mitochondrial DNA, this would translate into population-specificgenetic signatures only after an adequate amount of time has elapsed. In Europethere has evidently been sufficient time since divergence from a common ancestorfor population-specific genetic signatures to have evolved, a process that will havebeen accelerated by the extinction–recolonization dynamics inherent to a meta-population (see below).

Synthesis of genetic dataAlthough many populations of C. mucedo in northwest Europe are linked by geneflow, they are none the less genetically distinct. These results conform to theoreticalevidence that founder effects and genetic bottlenecks will result in population differentiation within a metapopulation (Wade & McCauley 1988). A variety ofevidence supports metapopulation dynamics in C. mucedo populations in Europe, including: apparent extinction and colonization events in UK populations(Okamura 1997b); historical records of multiple disappearances (up to some 30years) and reappearances of statoblasts in sediment cores from two UK sites (coresdated from 1740 and from 1963 to present) (R. Shaw et al., unpubl. data); large fluctuations in abundance within UK and Austrian populations (Wöss 1994;Vernon et al. 1996); infection by myxozoan and microsporidian parasites that maycause populations to fluctuate and occasionally go extinct (Okamura 1997a); andgenetic evidence of a bottleneck in a Scottish population based on microsatellitedata (Freeland et al. 2000a) and of genetic drift and founder effects in populationsanalysed by RAPD (random amplification of polymorphic DNA) in southern England (Hatton-Ellis et al. 1998). A metapopulation in Europe is further suggestedby microsatellite data which revealed similar within-population levels of genetic di-versity on both continents, whereas overall levels of genetic diversity were consider-ably lower in Europe. This conforms to the prediction that the effective size of ametapopulation should be smaller than the sum of its component subpopulations(Hedrick & Gilpin 1997).

Our results could be explained by sampling regimes. The European populationswere located along a transect that corresponds to an annual migratory route con-verged on by many waterfowl species (Fig. 10.4a) (Scott & Rose 1996).Waterfowl areobvious agents for long-distance dispersal of propagules of freshwater invertebrates(for review see Bilton et al. 2001), and sites located along a common annual migra-tory route may be expected to maintain a level of connectivity. Furthermore, traitssuch as small size, marginal hooks and spines (Fig. 10.3), and protective, chitinousvalves are likely to promote transport of C. mucedo statoblasts by waterfowl throughentanglement in feathers, or internally, if consumed. Indeed, moulted feathers withadherent statoblasts are often encountered around lake margins (B. Okamura, pers.comm.), while experiments demonstrating the viability of C. mucedo and other statoblasts ingested by ducks (Brown 1933; Charalambides et al., unpubl. data) and

205

Page 13: Chapter 10 Gene flow and the evolutionary ecology of passively

the presence of C. mucedo statoblasts in waterfowl digestive tracts and faeces (A.Green, unpubl. data) provide evidence for internal transport. In contrast, the sitessampled in North America were located along multiple annual migratory routes(Fig. 10.4b) and relatively low levels of connectivity among populations may be ex-plained by the divergent nature of routes and the locations of waterfowl stop-oversites across the region sampled. None the less, multiple haplotypes within NorthAmerican populations provide evidence for widespread postglacial dispersal, and itis possible that ongoing dispersal continues among populations that were not sampled.

A final consideration is the possibility of at least two cryptic Cristatella species inNorth America, as is suggested by the segregation of combined mtDNA and micro-satellite data into two distinct classes of genotypes (Freeland et al. 2000c) and the apparent existence of two genome sizes in North American C. mucedo (Potter 1979).Combining data from two species would artificially reduce our estimates of geneflow.

Evidence for dispersal in zooplankton taxaWe now review evidence for dispersal in several zooplankton groups by consideringthe results of population genetic and ecological studies.

Population genetic studies of zooplanktonCopepods, and especially cladocerans, are the most extensively studied freshwaterzooplankton taxa with respect to patterns of population genetics. Many species arecharacterized by endemic distributions and highly genetically differentiated popu-lations (Boileau & Hebert 1988, 1991; Crease et al. 1990; Boileau et al. 1992; Hebert& Wilson 1994; Hebert & Finston 1996; Taylor et al. 1998). In addition, the geneticdissimilarity of neighbouring populations reveals no pattern of isolation by distance(Hebert & Moran 1980; Hebert et al. 1989; Weider 1989; Hebert & Finston 1996;Vanoverbeke & De Meester 1997), and estimates of gene flow among populations are generally very low (Hebert & Moran 1980; Crease et al. 1990; Boileau et al. 1992).Recent studies of rotifers and ostracods demonstrate similar patterns of regional endemism, significant geographical structuring and reduced gene flow (Gómez et al. 2000; Schön et al. 2000). However, some cladoceran populations are separatedby only small genetic differences even over very large spatial scales (Innes 1990;Cerny& Hebert 1993; Hann 1995). These exceptions suggest that dispersal as a life-historytrait has attained varying levels of importance in different zooplankton species andpopulations. None the less, it is widely recognized that former views of broad speciesdistributions achieved through exceptional dispersal capabilities of zooplanktonmust now be tempered (see Schwenk et al. 2000 and references therein).

Cladocerans show phenotypic stasis associated with genetically divergent lin-eages, and many traditionally recognized cladoceran taxa have been revealed as anarray of cryptic species (Taylor et al. 1998; Schwenk et al. 2000). Such cryptic speciescomplexes presumably result from low levels of gene flow leading to genetic diver-gence and speciation. Speciation mechanisms include microevolution following

206 . . .

Page 14: Chapter 10 Gene flow and the evolutionary ecology of passively

range fragmentation of an ancestral species by physical barriers, disruptive selec-tion, and interspecific hybridization and introgression that results from reproduc-tive compatibility despite marked genetic divergence (for review see Schwenk et al.2000). The rapid evolution of mean resistance to nutritionally poor or toxiccyanobacteria by Daphnia galeata following eutrophication-driven changes in phy-toplankton assemblages provides further graphic evidence of significant and pre-cipitous genetic change within zooplankton populations (Hairston et al. 1999b), asdoes the evolution of body size in Daphnia in response to fluctuating selection within a season (Tessier et al. 1992), and genetic changes in a Daphnia population inresponse to changes in predation pressure over a 30-year period (Cousyn et al. 2001).

Although regular, ongoing gene flow is apparently rare amongst most zooplank-ton populations within regions, genetic evidence indicates that dispersal can be important in promoting range expansions from postglacial refugia, in limiting zoo-plankton at very broad spatial scales (Hebert & Hann 1986; Boileau & Hebert 1991;Stemberger 1995), and in introducing exotic species (Hairston et al. 1999a; Duffy et al. 2000) as further discussed below. As for bryozoans, genetic evidence impli-cates waterfowl as dispersal agents, with distributions of genetic lineages showingconcordance with annual migratory routes (Weider et al. 1996; Weider & Hobæk1997; Taylor et al. 1998).

Ecological studies of zooplanktonRecent ecological studies provide evidence that patterns of zooplankton dispersalcan play an important role in regulating community structure and function. Jenkinsand Buikema (1998) found highly variable patterns of colonization by multiple zoo-plankton species (rotifers, cladocerans, copepods and phantom midges) and subse-quent variation in community development in a series of environmentally similar,artificial ponds. Only 23% of zooplankton taxa from the regional species pool werepresent in all ponds after 1 year. Variation in zooplankton community compositionamong ponds was due to the behaviour of 75% of taxa in the regional species pool,with individual zooplankton species varying widely in their colonization successand timing. Overall, ponds differed in zooplankton community structure as measured by diversity-based (colonization and accrual curves, presence/absence,species richness) and population-based (density and biomass) data (Fig. 10.9). Thisstudy obtained evidence that, although dispersal is a rate-limiting process in deter-mining zooplankton community structure, community development reflects con-tingencies of colonization histories in local sites. Thus, historical events producepriority effects which can have a lasting influence on subsequent community struc-ture (see also Holland & Jenkins 1998).

Shurin (2000) conducted experimental studies of dispersal limitation in zoo-plankton by introducing rotifer and crustacean species from the region into estab-lished zooplankton communities in ponds and assessing the invasion resistance of these intact communities. Overall, success of invasion was low, as >91% of the introduced species immediately went extinct. However, some introduced speciesmanaged to colonize sites successfully, although they remained rare. Greater rates of

207

Page 15: Chapter 10 Gene flow and the evolutionary ecology of passively

208 . . .

Rotifera

7500

5000

2500

0

Copepoda

Cladocera

750

500

250

0

150

100

50

0

4

3

1

2

0

Chaoborus

F M A M J J A S O N D J F

Months

Num

ber

of o

rgan

ism

s l–1

Figure 10.9 Zooplankton taxa (phantom midges (Chaoborus), cladoceran, copepod, rotifer)density over time in a series of 12 experimental ponds whose colonization patterns weremonitored over a 1-year period. Each line represents a pond. (From Jenkins & Buikema1998, with permission.)

Page 16: Chapter 10 Gene flow and the evolutionary ecology of passively

209

invasion were observed when the abundance of native species was experimentallyreduced,providing evidence that interactions with resident species were responsiblefor excluding potential invaders. These results indicate a minor role for dispersallimitation in structuring pond zooplankton communities, and complement the priority effects identified by Jenkins and Buikema (1998) in showing that estab-lished residents can inhibit colonization.

In an examination of patterns revealed by 25 studies of regional and local speciesrichness in freshwater crustacean zooplankton, Shurin et al. (2000) came to the conclusion that the relative strengths of local and regional processes in determiningspecies richness depend on the regional scale studied. Dispersal limitation was con-cluded to be important at very large spatial scales, while local processes determinezooplankton diversity within a biogeographical region.

Synthesis of genetic and ecological investigations of zooplanktonThe above studies suggest that gene flow amongst zooplankton populations may behampered by the difficulty of invading established communities. Microevolutionwithin sites may further exacerbate problems of successful colonization followingdispersal if established residents are better adapted to exploit local conditions thanpotential colonists (for further discussion, see De Meester 1996). On the face of it,this is at odds with instances of invasions by exotic species (e.g. Berg & Garton 1994;Havel et al. 1995) and the colonization of new biogeographical regions by occa-sional long-distance dispersal events (Hebert & Hann 1986; Boileau & Hebert 1991;Stemberger 1995). This discrepancy may be explained, in some cases, if dispersal tonew biogeographical regions reflects range expansion from glacial refugia into un-saturated communities (e.g. Boileau & Hebert 1991; Weider & Hobæk 1997; Weideret al. 1999a, 1998b). In addition, invasion by exotics may often be explained bychanges in habitat conditions. For instance, invasion by pollution-tolerant, exoticDaphnia species followed habitat changes due to pollution (Hairston et al. 1999a;Duffy et al. 2000). In other cases, invasions may be temporary, as suggested by apparent postinvasion extinction in some sites (Havel et al. 1995). Finally, attri-butes of some invasive species may specifically promote invasion and exploitation of established communities. Such attributes include short generation times, beinghabitat and trophic generalists (Ehrlich 1986) and filling an empty niche (Johnson & Carlton 1996; see also Cohen, this volume).

While local adaptation may reconcile seemingly low levels of gene flow with re-peated observations of rapid colonization of new sites, a consideration that shouldbe addressed is the possibility that gene flow estimates may be unreliable (Whitlock& McCauley 1999). In contrast to many assumptions about gene flow calculations,population genetic structure seldom reflects a precise relationship between geneflow and genetic drift (Bohonak 1999). For example, estimates of gene flow may beparticularly biased if populations have arisen from recent (re)colonization, aprocess that requires dispersal, but which may not reveal evidence of gene flow because of stochastic processes associated with founder effects (e.g. Nurnberg &Harrison 1995). However, it is worth noting that despite the inherent difficulties in

Page 17: Chapter 10 Gene flow and the evolutionary ecology of passively

inferring gene flow from genetic data, gene flow estimates are seldom so over-whelmed by factors such as population history or disequilibrium as to be uninfor-mative (Bohonak 1999).

Of course practical issues may also influence detection of gene flow, for instance ifpopulations are characterized using relatively invariant markers such as allozymesor some mitochondrial sequence. In addition, many zooplankton populations areextremely large and inhabit numerous water bodies. As a result, for logistic reasons,sampled individuals may not provide a comprehensive assessment of populationstructure. We can expect that as new methods of genetic analyses evolve and furtherstudies are completed, our perception of gene flow among populations may con-tinue to change.

Temporal gene flow via propagule banksThere is accumulating evidence that dormant propagules produced by many fresh-water invertebrates are retained within sediments where they remain viable overprolonged periods of time and can thus act in the same way as plant ‘seed banks’,promoting temporal gene flow within local populations. Such resting egg bankshave been found in cladocerans (Caceres 1998),copepods (Hairston et al. 1995) androtifers (Gómez & Carvalho 2000). In addition, statoblast banks of C. mucedo havebeen inferred from population genetic studies which suggest repopulation from sta-toblasts released from sediments within a site, rather than from colonizers originat-ing elsewhere (Freeland et al. 2001). The extent to which statoblast banks in C.mucedo may counteract extinction events, thereby confounding interpretations ofmetapopulation dynamics, remains to be determined (see also Bullock et al., thisvolume). However, few statoblasts have remained viable beyond 6 years in labo-ratory studies, and most appear to be inviable after 2 years (Bushnell & Rao 1974;Mukai 1982;Wood 1991).The general pattern is for statoblast viability to decline withage. Criteria based on size and taxonomic similarity suggest a viability of 4–8 years forstatoblasts of C. mucedo. In contrast, the duration of viability of zooplankton eggs isin the order of decades to several hundreds of years (e.g. Hairston et al. 1995). It is tobe expected that the dynamics of local populations of C. mucedo will be influenced by a mixture of both spatial and short-term (<10 years) temporal gene flow.

Emergence from propagule banks may be particularly significant in years whenreproduction is poor. Thus, gains achieved during good years can compensate forlosses sustained in poor years, thereby providing a storage effect in temporally vary-ing environments (De Stasio 1990; Hairston et al. 1996). These repeated releases oflineages can also maintain local genetic variation, thereby influencing the rate ofevolution in response to switches in selection pressures (Hairston & De Stasio 1988;Ellner et al. 1999) which may be reflected in genetic shifts in propagule banks (Weider et al. 1997; Cousyn et al. 2001). The presence of heterozygote deficienciesmay reflect a temporal Wahlund effect if the resting propagule bank contains an admixture of propagules produced by temporally isolated populations differing intheir allele frequencies (Freeland et al. 2000a; Gómez & Carvalho 2000), and indeed

210 . . .

Page 18: Chapter 10 Gene flow and the evolutionary ecology of passively

such a temporal Wahlund effect may contribute to the high levels of FIS in C. mucedopopulations referred to earlier.

For both zooplankton and bryozoans, temporal gene flow provided by hatchingof propagules that have been retained in sediments may often be of greater impor-tance to population genetic structure than spatial gene flow. These taxa rely on pas-sive dispersal to achieve colonization of new sites, and such dispersal is likely to be arelatively rare event, contributing in part to the low to non-existent levels of geneflow that are typically observed (see earlier discussions). Periodic reintroduction ofdormant propagules from the sediment may be a much more reliable means ofachieving dispersal. Temporal gene flow may, therefore, play a pivotal role in boththe maintenance of genetic diversity and the avoidance of extinction in local popu-lations. In addition, in taxa such as C. mucedo that produce asexual dormant pro-pagules, temporal gene flow may promote the long-term persistence of cloneswithin sites. Finally, although the storage effect provided by propagule banks mayslow the rate of microevolution in response to fluctuating selection (Hairston & DeStasio 1988; Ellner et al. 1999), propagules banks may nevertheless promote localadaptation by enhancing the long-term persistence of populations within sites.

Dispersal, microevolution, life history and ecological patterns:lessons from bryozoans and zooplanktonPopulation genetic studies of freshwater invertebrates of lakes and ponds are begin-ning to provide a link between microevolution within populations and the resultantdynamics of those populations. Investigations of zooplankton taxa commonly reveal extensive local adaptation reflected in high levels of genetic differentiationamongst populations and cryptic speciation. Ecological studies of established andexperimental zooplankton communities show that colonization by new species is often precluded, despite dispersal, because communities are often saturated(Jenkins & Bukema 1998; Shurin 2000; Shurin et al. 2000). This combined informa-tion from ecological and genetic studies suggests that microevolutionary processeswithin sites determine the relative importance of regional versus local influences onzooplankton population and community structure. It also suggests that the appar-ent absence of metapopulation structures associated with zooplankton taxa can beexplained if local adaptation precludes ongoing gene flow between sites.

In contrast, genetic investigations of populations of C. mucedo reveal ongoinggene flow that maintains a metapopulation across northwest Europe. In thismetapopulation, the genetic differentiation which characterizes local populationsmay be greatly influenced by founder effects during (re)colonization, followed byinbreeding, genetic drift and temporal gene flow via statoblast banks. None the less,the identification of multilocus genotypes at multiple sites suggests the persistenceof general purpose genotypes that are adapted to the broad range of environmentalconditions present within this metapopulation, although the gradual erosion ofclonal diversity within a season is suggestive of local clonal selection (Hatton-Ellis1997; Freeland et al. 2001).

211

Page 19: Chapter 10 Gene flow and the evolutionary ecology of passively

North American populations of C. mucedo show intriguing differences from populations in northwest Europe. In particular, they reveal lower levels of gene flow,lack of evidence for a metapopulation structure, and the possibility of cryptic speci-ation. There are a number of potential explanations for the high levels of genetic dif-ferentiation among these North American populations. These explanations includesampling subpopulations from different metapopulations and processes that characterize zooplankton taxa such as infrequent spatial gene flow and local adapta-tion. Identifying the major determinants of genetic structure in North Americanpopulations will require further investigation.

It is notable that genetic and ecological data indicate a variable role of sexual re-production in bryozoan and zooplankton populations. C.mucedo undergoes a brief,early period of sexual activity which entails at least some outcrossing (J. Freeland,unpubl. data.), although a sexual phase appears to be foregone in some populationsat least in some years (Okamura 1997b). None of the C. mucedo populations thathave been characterized were in Hardy–Weinberg equilibrium, and marked ob-served heterozygosity deficiencies provide evidence for both inbreeding and theWahlund effect. In contrast, populations of many zooplankton taxa undergo sexualreproduction at the end of the season and generally conform to Hardy–Weinbergproportions (De Meester 1996; Gómez & Carvalho 2000). Since sex may play an im-portant role in contributing to local adaptation through the generation of geneticdiversity,there is an intriguing possibility that variation in the timing and expressionof sexual reproduction may explain variation in levels of population divergence andlocal adaptation. In zooplankton taxa, regular sexual production of resting stageswill generate significant genetic diversity in the population and hence will promoteadaptive change. In bryozoans, the asexual production of resting stages will at bestonly maintain extant levels of genetic diversity into the next season. Further factorsthat may depress adaptive change in bryozoans are inbreeding, irregular sex and theerosion of genetic diversity through the growing season by local selection prior tothe asexual production of overwintering statoblasts.

ConclusionsPopulation genetic and ecological studies of freshwater invertebrates are beginningto reveal how the frequency and scale of dispersal and gene flow relate to ecologicaland evolutionary patterns. Occasional, ongoing dispersal appears to maintain awidespread metapopulation structure that unites some bryozoan populations.However, in many zooplankton taxa colonization of new sites appears to be rela-tively infrequent and evolutionary divergence following postglacial expansion hasresulted in local adaptation and cryptic speciation. Further study of freshwater in-vertebrate taxa will help to unravel the specific influences of gene flow in space andtime, and may also clarify the conditions that promote dispersal and metapopula-tion structure versus population subdivision and accelerated adaptive change.

212 . . .

Page 20: Chapter 10 Gene flow and the evolutionary ecology of passively

213

ReferencesBerg, D.J. & Garton, D.W. (1994) Genetic differenti-

ation in North American and European popula-tions of the cladoceran Bythotrephes. Limnologyand Oceanography 39, 1503–1516.

Bilton, D., Freeland, J.R. & Okamura, B. (2001) Dis-persal in freshwater invertebrates. Annual Reviewof Ecology and Systematics 32, 159–181.

Bohonak,A.J. (1999) Dispersal, gene flow, and population structure. Quarterly Review of Biology74, 21–45.

Boileau, M.G. & Hebert, P.D.N. (1988) Genetic dif-ferentiation of fresh-water pond copepods atArctic sites. Hydrobiologia 167, 393–400.

Boileau, M.G. & Hebert, P.D.N. (1991) Genetic con-sequences of passive dispersal in pond-dwellingcopepods. Evolution 45, 721–733.

Boileau, M.G., Hebert, P.D.N. & Schwartz, S.S.(1992) Nonequilibrium gene frequency diver-gence: persistent founder effects in natural populations. Journal of Evolutionary Biology 5,25–39.

Brown, C.J.D. (1933) A limnological study of cer-tain fresh-water Polyzoa with special reference totheir statoblasts. Transactions of the American Microscopical Society 52, 271–314.

Bushnell, J.H. & Rao, K.S. (1974) Dormant or quies-cent stages and structures among the Ectoprocta:physical and chemical factors affecting viabilityand germination of statoblasts. Transactions ofthe American Microscopical Society 93, 524–543.

Caceres, C.E. (1998) Interspecific variation in theabundance, production, and emergence ofDaphnia diapausing eggs. Ecology 79, 1699–1710.

Cerny, M. & Hebert, P.D.N. (1993) Genetic diversityand breeding system variation in Daphnia puli-caria from North American lakes. Heredity 71,497–507.

Cornell, H.V. & Karlson, R.H. (1997) Local and regional processes as controls of species richness.In: Spatial Ecology: The Role of Space in PopulationDynamics and Interspecific Interactions (eds D.Tilman & P. Kareiva), pp. 250–268. PrincetonUniversity Press, Princton, NJ.

Cornell, H.V. & Lawton, J.H. (1992) Species interac-tions, local and regional processes, and limits tothe richness of ecological communities: a theo-retical perspective. Journal of Animal Ecology 61,1–12.

Cousyn, C., De Meester, L., Colbourne, J.K.,Bendonck, L.,Verschuren, D. & Volckaert, F.(2001) Rapid local adaptation of zooplankton behavior to changes in predation pressure in absence of neutral genetic changes. Proceedings of the National Academy of Sciences of the USA 98,625–626.

Crease, T.J., Lynch, M.M. & Spitze, K. (1990) Heir-archical analysis of population genetic variationin mitochondrial and nuclear genes of Daphniapulex. Molecular Biology and Evolution 7,444–458.

Darwin, C. (1859) On the Origin of Species by Meansof Natural Selection. John Murray, London.

De Meester, L. (1996) Local genetic differentiationand adaptation in freshwater zooplankton populations: patterns and processes. Ecoscience3, 385–399.

De Stasio Jr, B.T. (1990) The role of dormancy andemergence patterns in the dynamics of a fresh-water zooplankton community. Limnology andOceanography 35, 1070–1090.

Duffy, M.A., Perry, L.J., Kearns, C.M.,Weider, L.J. &Hairston Jr, N.G. (2000) Paleogenetic evidencefor a past invasion of Onondaga Lake, New York,by exotic Daphnia curvirostris using mtDNAfrom dormant eggs. Limnology and Oceanogra-phy 45, 1409–1414.

Ehrlich, P.R. (1986) Which animal will invade? In:Ecology of Biological Invasions of North Americaand Hawaii (eds H.A. Mooney & J.A. Drake), pp.79–95. Springer Verlag, New York.

Ellner, S.P., Hairston Jr, N.G., Kearns, C.M. & Babaï,D. (1999) The roles of fluctuating selection andlong-term diapause in microevolution of dia-pause timing in a freshwater copepod. Evolution53, 111–122.

Freeland, J.R., Noble, L.R. & Okamura, B. (2000a)Genetic consequences of the metapopulation biology of a facultatively sexual freshwater inver-tebrate. Journal of Evolutionary Biology 13,383–395.

Freeland, J.R., Noble, L.R. & Okamura, B. (2000b)Genetic diversity of North American populationsof Cristatella mucedo, inferred from microsatel-lite and mitochondrial DNA. Molecular Ecology 9,1375–1389.

Freeland, J.R., Romualdi, C. & Okamura, B. (2000c)Gene flow and genetic diversity: a comparison of

Page 21: Chapter 10 Gene flow and the evolutionary ecology of passively

214 . . .

freshwater bryozoan populations in Europe andNorth America. Heredity 85, 498–508.

Freeland, J.R., Rimmer,V.K. & Okamura, B. (2001)Genetic changes within freshwater bryozoanpopulations suggest temporal gene flow from statoblast banks. Limnology and Oceanography46, 1121–1129.

Gómez,A. & Carvalho, G.R. (2000) Sex, partheno-genesis and genetic structure of rotifers: micro-satellite analysis of contemporary and resting egg bank populations. Molecular Ecology 9,203–214.

Gómez,A., Carvalho, G.R. & Lunt, D.H. (2000) Phy-logeography and regional endemism of a pas-sively dispersing zooplankter: mitochondrialDNA variation in rotifer resting egg banks. Pro-ceedings of the Royal Society of London, Series B267, 2189–2197.

Goodman, S.J. (1997) R-st calc: a collection ofcomputer programs for calculating estimates ofgenetic differentiation from microsatellite dataand determining their significance. MolecularEcology 6, 881–885.

Hairston Jr, N.G. & De Stasio Jr, B.T. (1988) Rate ofevolution slowed by a dormant propagule pool.Nature 336, 239–242.

Hairston Jr, N.G.,Van Brunt, R.A. & Kearns, C.M.(1995) Age and survivorship of diapausing eggsin a sediment egg bank. Ecology 76, 1706–1711.

Hairston Jr, N.G., Ellner, S. & Kearns, C.M. (1996)Overlapping generations: the storage effect andthe maintenance of biotic diversity. In: Popula-tions Dynamics in Ecological Space and Time (edsO.E. Rhodes, R.K. Chesser & M.H. Smith), pp.109–145. Chicago University Press, Chicago.

Hairston Jr, N.G., Perry, L.J., Bohonak,A.J., Fellows,M.Q. & Kearns, C.M. (1999a) Population biologyof a failed invasion: paleolimnology of Daphniaexilis in upstate New York. Limnology andOceanography 44, 477–486.

Hairston Jr, N.G., Lampert,W., Cáceres, C.E. et al.(1999b) Rapid evolution revealed by dormanteggs. Nature 401, 446.

Hann, B.J. (1995) Genetic variation in Simocephalus(Anomopoda, Daphniidae) in North America:patterns and consequences. Hydrobiologia 307,9–14.

Hanski, I. (1989) Metapopulation dynamics: does ithelp to have more of the same? Trends in Ecologyand Evolution 66, 335–343.

Hanski, I. (1999) Metapopulation Ecology. OxfordUniversity Press, Oxford.

Hanski, I. & Simberloff, D. (1997) The metapopula-tion approach, its history, conceptual domain,and application to conservation. In: Metapopula-tion Biology. Ecology, Genetics, and Evolution (edsI. Hanski & M.E. Gilpin), pp. 5–26. AcademicPress, San Diego.

Harrison, S. & Taylor,A.D. (1997) Empirical evi-dence for metapopulation dynamics. In:Metapopulation Biology. Ecology, Genetics, andEvolution (eds I. Hanski & M.E. Gilpin), pp.27–42. Academic Press, San Diego.

Hatton-Ellis, T.W. (1997) The molecular ecology ofCristatella mucedo (Bryozoa: Phylactolaemata) inspace and time. PhD thesis, University of Bristol,Bristol.

Hatton-Ellis, T., Noble, L.R. & Okamura, B. (1998)Genetic variation in a freshwater bryozoan. I.Populations in the Thames Basin. Molecular Ecol-ogy 7, 1575–1585.

Havel, J.E., Mabee, M.R. & Jones, J.R. (1995) Inva-sion of the exotic cladoceran Daphnia lumholtziinto North American reservoirs. Canadian Journal of Fisheries and Aquatic Sciences 52, 151–160.

Hebert, P.D.N. & Finston, T.L. (1996) Genetic dif-ferentiation in Daphnia obtusa: a continental per-spective. Freshwater Biology 35, 311–321.

Hebert, P.D.N. & Hann, B.J. (1986) Patterns in thecomposition of Arctic tundra pond micro-crustacean communities. Canadian Journal ofFisheries and Aquatic Sciences 43, 1416–1425.

Hebert, P.D.N. & Moran, C. (1980) Enzyme variability in natural populations of Daphnia carinata King. Heredity 45, 313–321.

Hebert, P.D.N. & Wilson, C.C. (1994) Provincialismin plankton: endemism and allopatric speciationin Australian Daphnia. Evolution 48, 1333–1349.

Hebert, P.D.N., Beaton, M.J., Schwartz, S.S. & Stanton, D.J. (1989) Polyphyletic origins ofasexuality in Daphnia pulex. 1. Breeding systemvariation and levels of clonal diversity. Evolution43, 1004–1015.

Hedrick, P.W. & Gilpin, M.E. (1997) Genetic effec-tive size of a metapopulation. In: MetapopulationBiology (eds I. Hanski & M.E. Gilpin), pp.166–182. Academic Press, San Diego.

Holland, T.A. & Jenkins, D.G. (1998) Comparisonof processes regulating zooplankton assemblages

Page 22: Chapter 10 Gene flow and the evolutionary ecology of passively

215

in new freshwater pools. Hydrobiologia 387,207–214.

Innes, D.J. (1990) Geographic patterns of geneticdifferentiation among sexual populations ofDaphnia pulex. Canadian Journal of Zoology 69,995–1003.

Jenkins, D.G. & Buikema Jr,A.L. (1998) Do similarcommunities develop in similar sites? A test withzooplankton structure and function. EcologicalMonographs 68, 421–443.

Johnson, L.E. & Carlton, J.T. (1996) Post-establishment spread in large-scale invasions:dispersal mechanisms of the zebra mussel Dreissena polymorpha. Ecology 77, 1686–1690.

Levins, R. (1969) Some demographic andgenetic consequences of environmental heterogeneity for biological control. Bulletin ofthe Entomological Society of America 15, 237–240.

Levins, R. (1970) Extinction. Lecture Notes in Mathematical Life Sciences 2, 75–107.

Lynch, M. & Spitze, K. (1994) Evolutionary genetics of Daphnia. In: Ecological Genetics (ed.L.A. Real), pp. 109–128. Princeton UniversityPress, Princeton, NJ.

Menéndez, R. & Thomas, C.D. (2000) Metapopula-tion structure depends on spatial scale in thehost-specific moth Wheeleria spilodactylus(Lepidoptera: Pterophoridae). Journal of AnimalEcology 69, 935–951.

Mukai, H. (1982) Development of freshwater bry-ozoans (Phylactolaemata). In: Developmental Biology of Freshwater Invertebrates (eds F.W.Harrison & R.R. Cowden), pp. 535–576. Alan R.Liss, New York.

Nurnberg, B. & Harrison, R.G. (1995) Spatial popu-lation structure in the whirligig beetle Dineutusassimilis: evolutionary inferences based on mito-chondrial DNA and field data. Evolution 49,266–275.

Okamura, B. (1997a) Genetic similarity, parasitism,and metapopulation structure in a cyclicallyclonal bryozoan. In: Evolutionary Ecology ofFreshwater Animals (eds B. Streit, T. Stäedler & C.Lively), pp. 294–320. Birkhäuser Verlag, Basel,Switzerland.

Okamura, B. (1997b) The ecology of subdividedpopulations of a clonal freshwater bryozoan insouthern England. Archiv für Hydrobiologie 141,13–34.

Pennak, R.W. (1989) Freshwater Invertebrates of theUnited States. Protozoa to Mollusca, 3rd edn. JohnWiley & Sons, New York.

Potter, R. (1979) Bryozoan karyotypes and genomesizes. In: Advances in Bryozoology (eds G.P.Larwood & M.B. Abbott), pp. 11–32. AcademicPress, London.

Ricklefs, R.E. (1987) Community diversity: relativeroles of local and regional processes. Science 235,167–171.

Schön, I., Gandolfi,A., Di Masso, E. et al. (2000) Per-sistence of asexuality through mixed reproduc-tion in Eucypris virens (Crustacea, Ostracoda).Heredity 84, 161–169.

Schwenk, K., Posada, D. & Hebert, P.D.N. (2000)Molecular systematics of European Hyalodaph-nia: the role of contemporary hybridization inancient species. Proceedings of the Royal Society ofLondon, Series B 267, 1833–1842.

Scott, D.A. & Rose, P.M. (1996) Atlas of AnatidaePopulations in Africa and Western Eurasia.Wetlands International Publication No. 41.Wetlands International,Wageningen, the Netherlands.

Shurin, J.B. (2000) Dispersal limitation, invasion resistance, and the structure of pond zooplank-ton communities. Ecology 81, 3074–3086.

Shurin, J.B., Havel, J.E., Leibold, M.A. & Pinel-Alloul, B. (2000) Local and regional zooplanktonspecies richness: a scale-independent test for saturation. Ecology 81, 3062–3073.

Slatkin, M. (1987) Gene flow and the geographicstructure of natural populations. Science 236,787–792.

Srivastava, D.S. (1999) Using local–regional rich-ness plots to test for species saturation: pitfallsand potentials. Journal of Animal Ecology 68,1–16.

Stemberger, R.S. (1995) Pleistocene refuge areas andpost-glacial dispersal of copepods of the north-eastern United States. Canadian Journal of Fish-eries and Aquatic Sciences 52, 2197–2210.

Taylor, D.J., Hebert, P.D.N. & Colbourne, J.K. (1996)Phylogenetics and evolution of the Daphnialongispina group (Crustacea) based on 12S rDNAsequence and allozyme variation. Molecular Phy-logeny and Evolution 5, 495–510.

Taylor, D.J., Finston, T.L. & Hebert, P.D.N. (1998)Biogeography of a widespread freshwater crus-tacean: pseudocongruence and cryptic en-

Page 23: Chapter 10 Gene flow and the evolutionary ecology of passively

216 . . .

Weider, L.J., Lampert,W.,Wessels, M., Colbourne,J.K. & Limburg, P. (1997) Long-term geneticshifts in microcrustacean egg bank associatedwith anthropogenic changes in Lake Constanceecosystem. Proceedings of the Royal Society ofLondon, Series B 264, 1613–1618.

Weider, L.J., Hobæk,A., Colbourne, J.K., Crease,T.J., Defresne, F. & Hebert, P.D.N. (1999a) Holarctic phylogeography of an asexual speciescomplex: I. mtDNA variation in arctic Daphnia.Evolution 53, 777–792.

Weider, L.J., Hobæk,A., Hebert, P.D.N. & Crease,T.J. (1999b) Holarctic phylogeography of anasexual species complex: II. Allozymic variation in arctic Daphnia. Molecular Ecology 8,1–13.

Wetzel, R.G. (1975) Limnology. W.B. Saunders,Philadelphia.

Whitlock, M.C. & McCauley, D.E. (1999) Indirectmeasures of gene flow and migration: F-st π1/(4Nm + 1). Heredity 82, 117–125.

Williams, D.D. (1987) The Ecology of TemporaryWaters. Croom Helm, London.

Wood, T.S. (1991) Bryozoans. In: Ecology and Clas-sification of North American Freshwater Inverte-brates (eds J.H. Thorp & A.P. Covich), pp.481–499. Academic Press, New York.

Wöss, E.R. (1994) Seasonal fluctuations ofbryozoan populations in five water bodies withspecial emphasis on the life cycle of Plumatellafungosa (Pallas). In: Biology and Palaeobiology ofBryozoans (eds P.J. Hayward, J.S. Ryland & P.D. Taylor), pp. 211–214. Olsen & Olsen,Fredensborg, Denmark.

demism in the North American Daphnia laeviscomplex. Evolution 52, 1648–1670.

Tessier, S.J.,Young,A. & Leibold, M.A. (1992)Population dynamics and body size selection inDaphnia. Limnology and Oceanography 37, 1–13.

Vanoverbeke, J. & De Meester, L. (1997) Among-populational genetic differentiation in the cycleparthenogen Daphnia magna (Crustacea,Anomopoda) and its relation to geographic dis-tance and clonal diversity. Hydrobiologia 360,135–142.

Vernon, J.G., Okamura, B., Jones, C.S. & Noble, L.R.(1996) Temporal patterns of clonality and para-sitism in a population of freshwater bryozoans.Proceedings of the Royal Society of London, Series B263, 1313–1318.

Wade, M.J. & McCauley, D.E. (1988) Extinction andrecolonization: their effect on the genetic differ-entiation of local populations. Evolution 42,995–1005.

Weider, L.J. (1989) Population genetics of Polyphe-mus pediculus (Cladocera: Polyphemidae).Heredity 62, 1–10.

Weider, L.J. & Hobæk,A. (1997) Postglacial dispersal, glacial refugia, and clonal structure in Russian/Siberian populations of the arctic Daphnia pulex complex. Heredity 78, 363–372.

Weider, L.J., Hobæk,A., Crease, T.J. & Stibor, H.(1996) Molecular characterization of clonal population structure and biogeography of arcticapomictic Daphnia from Greenland and Iceland.Molecular Ecology 5, 107–118.