25
Behavioural, population, and genetic processes affecting metapopulation dynamics of the Glanville fritillary butterfly Alia Sarhan Department of Biological and Environmental Sciences University of Helsinki Finland Academic dissertation To be presented, with permission of the Faculty of Biosciences of the University of Helsinki, for public criticism in theAuditorium 3 in Viikki Building B (Latokartanonkaari 7) October 20 th 2006 at 12 oclock noon Helsinki 2006

Behavioural, population, and genetic processes affecting

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Behavioural, population, and genetic processes affecting

Behavioural, population, and genetic processes affectingmetapopulation dynamics of the Glanville fritillary

butterfly

Alia Sarhan

Department of Biological and Environmental SciencesUniversity of Helsinki

Finland

Academic dissertation

To be presented, with permission of the Faculty of Biosciences of theUniversity of Helsinki, for public criticism in the Auditorium 3 in Viikki Building B

(Latokartanonkaari 7) October 20th 2006 at 12 o clock noon

Helsinki 2006

Page 2: Behavioural, population, and genetic processes affecting

ã Alia Sarhan (chapter 0)ã Blackwell Publishing (chapter I)ã Authors (chapters II-V)

Author s address:Department of Biological and Environmental SciencesP.O. Box 65 (Viikinkaari 1)FI-00014 University of HelsinkiFinland

E-mail: [email protected]

ISBN 952-92-1091-4 (Paperback)ISBN 952-10-3441-6 (PDF, http://ethesis.helsinki.fi)

Layout: Camilla EkbladCover photo: Nina Lehtosalo

YliopistopainoHelsinki 2006

Page 3: Behavioural, population, and genetic processes affecting

Behavioural, population, and genetic processes affectingmetapopulation dynamics of the Glanville fritillary butterfly

ALIA SARHAN

The thesis is based on the following articles, which are referred in the text by their Romannumerals:

I Sarhan, A. 2006. Isolation and characterization of five microsatellite loci in theGlanville fritillary butterfly (Melitaea cinxia). Mol. Ecol.Notes 6:163-164.

II Sarhan, A., Ovaskainen, O., and Hanski, I. Size and composition of foundingpropagules colonizing empty habitat patches in a butterfly metapopulation(submitted)

III Sarhan, A. and Haikola, S. Inbreeding depression following establishment ofnew populations in a natural metapopulation of the Glanville fritillary butterfly(Melitaea cinxia) (submitted)

IV Sarhan, A. Protandry in the Glanville fritillary butterfly (Melitaea cinxia): earlyemergence of males is adaptive (submitted)

V Sarhan, A. and Kokko, H. Multiple mating in the Glanville fritillary butterfly: acase of within-generation bet-hedging? (Evolution, accepted)

Page 4: Behavioural, population, and genetic processes affecting

Contributions

AS Alia Sarhan, HK Hanna Kokko, IH Ilkka Hanski, JP Jodie Painter, OO Otso Ovaskainen, SHSari Haikola, MS Marjo Saastamoinen

I II III IV V

Original idea AS IH, AS AS AS AS, HK

Study design AS, JP IH, AS SH IH, MS IH, MS

Data gathering AS AS SH, AS MS MS

Analyses AS AS, OO AS, SH AS AS, HK

Manuscript preparation AS AS AS, SH AS AS, HK

Supervised by Prof. Ilkka HanskiUniversity of HelsinkiFinland

Review by Prof. Rauno Alatalo, University of Jyväskylä, FinlandProf. Pekka Pamilo, University of Oulu, Finland

Examined by Prof. Lukas Keller, University of Zürich-Irchel, Switzerland

Page 5: Behavioural, population, and genetic processes affecting

CONTENTS

0 Summary

Introduction 7The Glanville fritillary metapopulation a model system in population biology 7

Genetic variation in metapopulations 8Methods: microsatellite markers 10

Results and discussion 12Colonizations 12Inbreeding and local extinctions 13Life history evolution in metapopulations 14

Conclusions 16Acknowledgements 19References 19

I Isolation and characterization of five microsatellite loci in the Glanville fritillarybutterfly (Melitaea cinxia)

II Size and composition of founding propagules colonizing empty habitat patchesin a butterfly metapopulation

III Inbreeding depression following establishment of new populations in a naturalmetapopulation of the Glanville fritillary butterfly (Melitaea cinxia)

IV Protandry in the Glanville fritillary butterfly (Melitaea cinxia): early emergence ofmales is adaptive

V Multiple mating in the Glanville fritillary butterfly: a case of within-generationbet-hedging?

Page 6: Behavioural, population, and genetic processes affecting
Page 7: Behavioural, population, and genetic processes affecting

Introduction

The Glanville fritillary metapopulation -a model system in population biology

The Glanville fritillary metapopulation in theÅland Islands exhibits classic metapopulationstructure and dynamics (Hanski 1999; Hanski etal. 1995), and it has become a widely recognizedmodel system for understanding the rules ofpersistence of species living in fragmentedlandscapes, and for population biology in general(Ehrlich and Hanski 2004). The metapopulationoccupies a large network of about 4,000 discretesmall meadows within an area of 50 by 70 km, outof which about 500 meadows are occupied by abutterfly population at a time (Hanski 1999;Nieminen et al. 2004). The meadows are clusteredinto semi-independent patch networks thatvary in the number of patches, patch sizes andconnectivities. Annual surveys are conducted inthe autumn, when all habitat patches aresurveyed for the presence or absence of localpopulations and the number of larval groups iscounted. The butterfly has gregarious larvae thatspin a conspicuous winter nest, a feature thatgreatly facilitates annual surveys (Nieminen etal. 2004). A colonization event is recorded whenone or more larval groups are found in a habitatpatch that was unoccupied in the previousautumn. Similarly, an extinction event is recordedwhen no larval group is found in a habitat patchthat was occupied in the previous autumn.Though the larval groups are quite conspicuousand extra care is taken to record the presence orabsence of populations as carefully as possible,Nieminen et al. (2004) have estimated that theprobability of missing an existing local populationis from 0.10 to 0.15.

The metapopulation is characterized by a veryhigh turnover rate of populations, close to 50%.Local populations are on average very small (4.3family groups on average of mostly full-siblarvae), and new populations are frequentlyestablished by a single mated female (Hanski1999; Hanski et al. 1995).

In Åland, adults fly in June to early July. Femalesare believed to usually mate only once in theirlifetime in the field (Kuussaari 1998), althoughthey were often found to regularly mate with twoor even three males in an outdoor cage withrelatively high population densities (Hanski et al.2006). Females lay eggs in clusters of 100-200 ontheir host plants Plantago lanceolata or Veronicaspicata, and the larvae hatch after two to threeweeks. Because females only oviposit on the hostplants, ovipositions can be easily monitored inexperimental setups. Larvae overwinter in a winternest that they spin at the end of August. Larvaestart feeding again in the spring, and they pupatein May. There is a pronounced pattern ofprotandry, with males hatching on average 2-3days earlier than females.

An important cause of mortality in naturalpopulations of the Glanville fritillary is overwintermortality: 20% of entire larval groups die duringthe winter. Overwinter mortality is known to bedependent on larval group size, small groups ofless than 20 larvae having a very small chance ofsurviving (Kuussaari 1998; Nieminen et al. 2001).

The long-term empirical study of the Glanvillefritillary has stimulated new modelling approachesin metapopulation theory by providing abundant,high-quality data to test model assumptions andpredictions; and the study has thereby facilitatedthe development of metapopulation theory to beapplicable to real metapopulations (Ehrlich and

Summary

ALIA SARHAN

Metapopulation Research Group, Department of Biological and Environmental Sciences, PO Box65, FI 00014, University of Helsinki, Finland

7

Page 8: Behavioural, population, and genetic processes affecting

Hanski 2004). The long-term project has alsohelped integration of ecological, genetic andevolutionary processes. Because ecologicalphenomena are variable in space and time,meaningful integration of ecological and geneticresearch on natural populations can only beachieved in model systems for which there is large-scale and long-term demographic informationavailable. For instance, the extensive databaseon population sizes and other ecological factorsin several hundreds local populations for manyyears allowed Saccheri et al. (1998) to control forthe influence of ecological factors on extinctionrisk and to show that inbreeding depressionincreases the risk of extinction of small localpopulations. This demonstration would not havebeen possible if the research had not beenconducted in the model system context. Anotherexample of integration of ecological andevolutionary studies is the finding that extinction-colonization dynamics play a role in the evolutionof host plant use (Hanski and Heino 2003).Research on the Glanville fritillary has alsocontributed much empirical information andstimulated novel modelling work on the evolutionof migration rate in metapopulations. One keyresult is that colonizations select for increasedmigration rate in metapopulations with fastturnover rate (Hanski et al. 2002).

Although the Glanville fritillary system is verywell known in many respects, several processesthat are crucial for understanding the genetics ofmetapopulations are still unknown. For instance,concerning the colonization of empty habitatpatches, it is known that females are able to layseveral egg clutches. The number and relatednessof females colonizing new populations is thereforeunknown when there is more than one egg clutch.Additionally, although it is assumed that femalesmate in their natal population prior to dispersal,this has not empirically been investigated. If onlyone or two females establish new populations, itis more likely that the populations will suffer frominbreeding depression. Inbreeding depression hasbeen shown to occur in the laboratory for thisspecies (Haikola et al. 2001), but never in naturalpopulations. And although a correlation betweenlow heterozygosity of genetic markers and highextinction risk has been established (Saccheri et

al. 1998), the exact mechanism leading frominbreeding depression to population extinctionhas not been elucidated. Moreover, althoughfemales are assumed to generally mate only once,a large proportion of females has been observedto mate up to three times in a large outdoor cageexperiment that was conducted in nearly naturalconditions (Hanski et al. 2006). This largepopulation cage experiment also providesinformation on reproductive success ofindividuals (Hanski et al. 2006), and thereby allowsinvestigating questions related to the evolutionof life-history traits and various behaviouralstrategies in greater detail than previously.

The aims of this thesis work were to elucidatesome of the behavioural, population, and geneticprocesses that have remained poorly studied andwhich may greatly influence metapopulationdynamics. More specifically, I investigatedpatterns of colonizations in terms of number andrelatedness of founders, and studied inbreedingdepression resulting from populationestablishment by only one female for populationsof different age. Additionally, I analyzed spermprecedence patterns in multiply mated females toaddress the evolution of polyandry in ametapopulation context. I also used thatinformation to determine the reproductive successof males and investigate how adaptive isprotandry.

Genetic variation in metapopulations

Genetic variation is most easily perceived asadaptive variation in morphology, behaviour andphysiology. Beyond that, neutral genetic variationis a measure of a population s evolutionarypotential, the raw material for evolution thatenables organisms to adapt to changingconditions (Hedrick and Miller 1992). Neutralgenetic variation is often measured asheterozygosity of neutral markers or as theadditive genetic variance that underlies characterssuch as life-history traits and morphology. Theamount of genetic variation present in apopulation is shaped by selection, breedingstructure, genetic drift, gene flow and mutation,all processes that depend on the effectivepopulation size. Factors that decrease the

8

Page 9: Behavioural, population, and genetic processes affecting

effective population size include small censuspopulation size, variance in reproductive success,population bottlenecks, non-random mating,unequal sex ratios, variable population size andpolygynous mating systems (Gaggiotti 2003;Wang and Caballero 1999). The smaller theeffective population size, the more importantgenetic drift becomes, the change in allelefrequency from one generation to the next causedby random sampling of alleles. The reduction infitness caused by drift changing allele frequenciesaway from those favoured by selection isexpressed as drift load (Crow 1993). Genetic driftand mating among close relatives due to restrictedmating opportunities can lead to inbreedingdepression, a reduction in fitness due to theincreased expression of the genetic load, as wellas reduced opportunities to expressoverdominance (Keller and Waller 2002; Mitton1993). Inbreeding can adversely affectdevelopment, growth, survival, and fecundity(Charlesworth and Charlesworth 1987) and it hasbeen extensively studied in domestic animals andin captive populations (Allendorf and Leary 1986;Crnokrak and Roff 1999; Falconer and Mackay1996), and more recently also in the wild, mainlythrough intensive long-term studies with deeppedigree knowledge (Coltman et al. 1999; Coulsonet al. 1998; Keller et al. 1994).

Frequency-dependent selection, spatiallyheterogeneous selection and overdominance allresult in balancing selection and could act tomaintain genetic variation (Whitlock 2004). Instructured populations, reduced heterozygositycauses a reduction in the amount of variationmaintained by overdominance and results in amore pronounced segregation load (Whitlock2002). On the other hand, epistatic interactionsbetween loci can cause different alleles to befavoured locally and maintain genetic diversitythrough heterogeneous selection (Whitlock andBarton 1997).

The effects of population structure on themaintenance of genetic diversity are notstraightforward and depend on the variance infitness and degree of genetic differentiationamong local populations (Whitlock and Barton1997). However, it is well established that

recurrent local extinctions and recolonizations,by increasing the variance in reproductivesuccess among local populations and thusdecreasing the effective population size, cause amajor reduction in the genetic diversity maintainedwithin local populations and in the totalmetapopulation (Pannell and Charlesworth 1999;Whitlock and Barton 1997).

Population bottlenecks are common inmetapopulations with frequent extinctions andrecolonizations, especially when newpopulations are established by only few founders.As the maximum fraction of genetic variation lostduring a bottleneck is a function of the populationgrowth rate (Nei et al. 1975), populations thatrecover quickly after a bottleneck lose littlegenetic variation even if the population wasreduced to a few individuals only. However,although such bottlenecks may not have a largeeffect on heterozygosity, they will have a strongimpact on allelic diversity and thereby constitutea long term threat to population viability byreducing its evolutionary potential (Allendorf1986). The loss of genetic variation has the effectof decreasing the additive genetic variance,whereas when there are gene interactions thedifferentiation of populations will tend to causethe additive genetic variance to increase. Thus,whether the additive genetic variance will increaseor decrease after a bottleneck in structuredpopulations depends on the relative magnitudeof these two effects (Goodnight 2004).

Most new mutations are deleterious, and theyare eliminated by selection in large populations,but in small populations they can become fixed ifS< 1/Ne (Maruyama 1970). With hard selection,population structure always increases theresponse to selection, especially for nearlyrecessive alleles that can be expressed morestrongly in structured populations due to theexcess homozygosity. With soft selection,however, individuals are competing locally forresources, and as the relatedness betweencompeting individuals is increased withpopulation structure, the response to selectioncan be lowered relative to a non-structuredpopulation (Whitlock 2002). Although selectionis often more effective in structured populations,

9

Page 10: Behavioural, population, and genetic processes affecting

this effect is usually counterbalanced by astrongly decreased effective population size, sothat the rate of accumulation of deleterious allelestends to be strongly increased in structuredpopulations (Higgins and Lynch 2001; Whitlock2003). Thereby, slightly deleterious alleles tendto accumulate over time, reducing individualfitness and contributing to an increasing mutationload. Also the probability of fixation of beneficialalleles tends to be much reduced with populationstructure, as a result of the reduced effectivepopulation size. It can however be increased forsome loci, especially for nearly recessive allelesthat can be expressed more strongly in structuredpopulations because of increased homozygosity.Additionally, the time taken for fixation of newalleles is also affected by population structure(Whitlock 2003).

Migration will tend to decrease the populationmean fitness and hence reduce the level of localadaptation because migrants are likely to be poorlyadapted to local conditions. This will result inmigration load, which becomes particularlyimportant when there are substantial differencesin the selection coefficients among populations,and can even set range limits of species(Kirkpatrick and Barton 1997; Mayr 1963).Migration load is lowered with lower migrationrate, whereas drift load, segregation load and localdrift load tend to increase with reduced migrationrate. Whether population structure increases ordecreases mean fitness on average depends on alarge number of circumstances and on thespecifics of the species (Whitlock 2004).

Methods: microsatellite markers

Due to advances in laboratory techniques andmethods of data analysis, genetic approaches foraddressing ecological questions are becomingincreasingly efficient and powerful (Selkoe andToonen 2006). Microsatellites are one of the mostpopular marker type for these studies, becausethey are numerous, highly variable, welldistributed through the genome, consistentlyscorable and comparable, and easy to use(Schlötterer 2004). Microsatellites are tandemrepeats of 1-6 nucleotides that are found at high

frequency in the nuclear genomes of most taxaand typically vary in length between 5 and 40repeats. In molecular genetic studies mostlydinucleotides, trinucleotides and tetranucleotidesare used (Li et al. 2002). Microsatellite sequencesmutate frequently by slippage and proofreadingerrors during DNA replication, resulting primarilyin a change in the number of repeats (Eisen 1999).Because alleles differ in the number of repeatsand thus differ in length, they can bedistinguished by high resolution gelelectrophoresis, a much more rapid and cost-efficient method than sequencing. Primers aredesigned to bind to the generally conservedflanking regions of a microsatellite locus foramplification with polymerase chain reaction(PCR). PCR-based technology allows using evendegraded DNA for the amplification and makesnon-invasive sampling possible, as DNA can beextracted from e.g. hair or faecal samples. Becauseof their high mutation rate, on average 10-4

mutations per locus per generation,microsatellites exhibit high levels of allelicdiversity (Schlötterer 2000).

There are however several drawbacks tomicrosatellite markers. There are some taxa forwhich marker isolation is still fraught withconsiderable failure rate (Selkoe and Toonen 2006),such as some marine inverterbrate (Cruz et al.2005; Hedgecock et al. 2004), lepidopterans(Zhang 2004) and birds (Primmer et al. 1997). Inaddition, the mutational mechanisms ofmicrosatellites remain unclear. In the infinite allelemodel (IAM) every mutation event creates a newallele, whereas in the stepwise mutational model(SMM) that is specific to microsatellites, one ormore repeat units is added or subtracted at aconstant rate. Even though in the few modelorganisms studied microasatellites seem to followmainly the SMM model (Eisen 1999; Ellegren2004), metrics using the SMM tend to be highlysensitive to violations of that model, and thusmetrics using the IAM seem to be more robustand reliable (Landry et al. 2002). Most analysesare however insensitive to mutational mechanism(Neigel 1997). Another drawback of microsatellitesis that because of the high mutation rate, allelesthat do not differ in length are not necessarilyidentical by descent, mainly because of back-

10

Page 11: Behavioural, population, and genetic processes affecting

mutations and convergence. This can howeverbe accounted for in analyses (Estoup andCornuet 1999; Slatkin 1995), and appears to beproblematic only when comparing verygenealogically distant groups (Angers et al. 2000).Microsatellite markers should also be tested forgametic disequilibrium, selective neutrality andMendelian inheritance (Selkoe and Toonen 2006).Violation of Mendelian inheritance is mainly dueto null alleles (Selkoe and Toonen 2006).

Problems in the amplification and scoring of allelespresent the main drawback of usingmicrosatellites (Selkoe and Toonen 2006). Sourcesof error include misinterpreting an artefact peakas a true allele, incorrect interpretation of stutterpattern, and null alleles (Selkoe and Toonen 2006).Proper sample preservation is particularlyimportant, as it can reduce substantially technicaldifficulties with amplification (Dawson et al. 1998).It is also important to always include a positivecontrol (Delmotte et al. 2001). Error rate can becalculated by repeating marker amplification andscoring in a random subset (Hoffman and Amos2005).

Null alleles are a very important source of error,particularly in certain taxa, like Lepidoptera(Meglecz and Solignac 1998). They can bedetected as heterozygote deficit when testing forHardy-Weinberg equilibrium, or if someindividuals repeatedly fail to amplify at just onelocus. Other possible causes of heterozygotedeficit include inbreeding, population structureand the Wahlhund effect (Nielsen et al. 2003).Biological causes of heterozygote deficit shouldnevertheless affect all loci similarly, whereas nullalleles affect only single loci. Statisticalapproaches to identify null alleles are alsoavailable (Van Oosterhout et al. 2004). Anotherway to detect null alleles is to examine patterns ofinheritance in a pedigree (Paetkau and Strobeck1995). Larger allele dropout is another way thatalleles can be missed, when the longer allele in aheterozygote does not amplify as well as theshorter one, and is too faint to be detected in thescoring process (Wattier et al. 1998). Redesigningprimers or adjusting PCR conditions can oftenameliorate null allele problems (Callen et al. 1993).

The development of microsatellites markers forLepidoptera has proven extremely difficult(Meglecz et al. 2004; Meglecz and Solignac 1998;Zhang 2004), apparently because microsatellitesequences with almost identical flanking regionsexist in multiple copies in the genome oflepidopteran species (Meglecz et al. 2004; Zhang2004). This is widely supported by the fact thatfrom all published microsatellite studies inLepidoptera no more than five polymorphic lociwere isolated per genomic library in 80% of thecases (Zhang 2004). In the first chapter of mythesis, I describe the isolation andcharacterization of five microsatellite markers forthe Glanville fritillary butterfly. All of them werecharacterized by very broad size ranges and highdegree of polymorphism, which is typical forLepidopera (Flanagan et al. 2002; Harper et al.2000; Keyghobadi et al. 1999; Meglecz andSolignac 1998). Highly variable microsatelliteshave distinct advantages and drawbacks (Selkoeand Toonen 2006). Genotype scoring errors mayoccur due to increased large allele dropout(Buchan et al. 2005) and increased stutter (Hoffmanand Amos 2005), and they have high rates ofhomoplasy that can introduce bias into allelefrequency estimates, dampen estimates of Fstvalues, and lead to substantial inflation of geneflow estimates (Gaggiotti et al. 1999; Slatkin 1995).On the other hand, highly variable loci haveincrease power to distinguish close relatives forparentage (Queller et al. 1993). In the followingchapters I have used these microsatellites forparentage identification when measuring thelifetime reproductive success of males,determining the sperm-precedence pattern inmultiply-mated females, and estimating thenumber of females that established new localpopulations in the metapopulation. I also usedmicrosatellites to estimate the relatedness ofindividuals.

Using microsatellite markers to estimateinbreeding has been heavily crtiticized (Pemberton2004; Slate et al. 2004) based on a study comparingmicrosatellite-derived estimates of heterozygosityto the heterozygosity obtained from pedigrees insheep (Slate et al. 2004). That study has notinvestigated other microsatellite-based measures,like internal relatedness, but suggests that such

11

Page 12: Behavioural, population, and genetic processes affecting

measures might be promising alternatives, as theyincorporate population allele frequencies (Slateet al. 2004). One reason why microsatellite derivedestimates of heterozygosity seemed to performso poorly in that study might have been the verylow variance in expected inbreeding coefficient,which made it very difficult for marker-basedestimates of heterozygosity to capture differencesbetween individuals (Slate et al. 2004). Becausemicrosatellites do not seem to be very suited toestimate inbreeding coefficients of individuals, Ionly used them to estimate the average inbreedingof local populations and the relatedness betweenindividuals. In these roles microsatellites seemedto perform well, and the estimates were in linewith what was expected based on the knowndemographic history of populations.

Results and discussion

Colonizations

The number and relatedness of individualscolonizing a new population are key parametersin determining the amount of genetic variation inlocal populations and in the entire metapopulationas well as the degree of differentiation amonglocal populations (Pannell and Charlesworth1999). Two extreme colonization models have beeninvestigated in the theoretical literature: in thepropagule pool model colonizers come from thesame local population whereas in the migrant poolmodel colonizers represent a subsample from theentire metapopulation (Pannell and Charlesworth1999; Slatkin 1977; Wade and McCauley 1988).Although the genetic consequences of differentcolonization models have been investigatedtheoretically, the colonization pattern has rarelybeen described in real metapopulations (Gaggiottiet al. 1999). The second chapter of my thesis aimsat determining how many females establish newlocal populations, and whether females mate intheir natal patch before dispersal. Additionally,in this study, which I did together with OtsoOvaskainen and Ilkka Hanski, the genetic resultsare compared with the results of an advanceddispersal model parameterized for the Glanvillefritillary. The model can be used to assess theexpected number of females colonizing new

populations given the sizes and spatial locationsof local populations in the previous generation.

The dispersal model captured well the generalpattern of colonizations. The high mortality duringthe early life instars (Nieminen et al. 2004) is onereason why there were rarely several larval groupsfrom the same female in newly-establishedpopulations in the empirical data. Larval mortalityis negatively correlated with larval group size(Kuussaari 1998) and it should therefore beparticularly high for later clutches, as these tendto be small (Wahlberg 1995). Because oursampling was done in the autumn, all larval groupsthat had died before that were missed. Anotherreason why females did not lay several eggclutches in the same local populations is that theymight not stay long enough in one patch to doso. The interval between ovipositions is usuallytwo days, but it can be up to two weeks if weatherconditions are unfavourable (Hanski, pers. obs.).

The genetic results show that butterfliesestablishing new populations were more relatedto their mate than to other parent butterflies inthat population, implying that females mate intheir natal local population prior to dispersal.These results confirm observations that femalesusually mate shortly after eclosure (Boggs andNieminen 2004). Half of the colonizationsconsisted of only one larval group, indicatingthat these populations were established by onlyone mated female. This situation is an extremeexample of propagule pool model with only tworelated founders and should result in very lowgenetic diversity in the newly foundedpopulations. These populations are also likely tosuffer from inbreeding depression, a subject Ihave investigated in the third chapter of this thesis.On the other hand, when there were several larvalgroups, the genetic results show that they weremostly established by several unrelated females.Because the females have nevertheless mated intheir natal patch, this situation represents anintermediate one between the propagule poolmodel and the migrant pool model, with severalunrelated groups of two related founders (afemale and her mate) establishing a newpopulation. In subsequent generations, matingsamong the offspring of these unrelated pairs

12

Page 13: Behavioural, population, and genetic processes affecting

should result in higher fitness because ofheterosis, a subject addressed in the third chapterof this thesis.

Inbreeding and local extinction

The Glanville fritillary butterfly metapopulationin the Åland Islands remains to date one of theonly two systems in which a correlation betweeninbreeding (measured as low heterozygosity atgenetic markers) and population extinction riskhas been convincingly demonstrated (Frankham2005; Saccheri et al. 1998). Although there is ampleevidence for inbreeding depression in this speciesin the laboratory (Haikola et al. 2001; Nieminen etal. 2001), the extent of inbreeding depression hasnot yet been quantified in natural populations. Inthe third chapter of this thesis, in a study I didwith Sari Haikola, I chose to concentrate on newly-established populations that were founded byonly one female. From the result of the previouschapter it is evident that this is a very commonsituation in the metapopulation, as about half ofthe new populations had been established by asingle female. Moreover, the fact that females werefound to mate in their natal patch prior to dispersalmakes these populations particularly prone tosuffer from inbreeding depression in thegenerations following population establishment.The demographic knowledge about populationage and population size that is available for manyyears allowed us to compare inbreedingdepression in the generation immediatelyfollowing population establishment ( newpopulations ) to that in population that hadremained small and isolated for severalgenerations after establishment ( oldpopulations ). Because in small populations allindividuals might suffer of inbreeding depressiondue to genetic drift, the only way to measure thedeleterious effects of inbreeding on fitness is bycrossing individuals between populations andobserving the fitness of their progeny (Hedrickand Kalinowski 2000; Keller and Waller 2002). Ametapopulation context is ideal for this kind ofstudy, because crossing individuals from differentpopulations corresponds to what happensnaturally when a migrant mates with residents(Ebert et al. 2002; Glemin et al. 2003; Haag et al.2002; Keller and Waller 2002; Richards 2000).

Our study design consisted of conductingmatings within and among new and oldpopulations, and to measure offspring fitness asegg-hatching rate and larval survival to diapause.Additionally, inbreeding levels in new and oldpopulations were estimated as averagepopulation heterozygosity at genetic markers, andpairwise relatedness between the two parentalindividuals was also computed. The metric I usedwas internal relatedness, a measure that takesinto account allele frequencies in the populationand that has been shown to be a much betterestimate of inbreeding than previously usedmetrics (Amos et al. 2001). Populations that hadbeen old and isolated for several years hadstrongly reduced heterozygosity levels comparedto newly-founded populations, indicating rapidloss of genetic variation in these populations.Offspring fitness was strongly reduced in matingswithin old population, with a decrease in bothegg-hatching rate and larval survival to diapause.Most importantly, as the reduction in egg-hatchingrate and larval survival cumulatively contributeto reduced larval group size, these effects arefurther amplified by increased larval-groupdependent overwinter mortality. Parentalrelatedness was found to be an important factorin determining offspring fitness, which declinedsharply with increasing parental relatedness. Thisstudy provides an important link between theresults of Saccheri et al (1998) of an increase inextinction risk in populations with low markerheterozygosity and laboratory results describinginbreeding depression in this species (Haikola etal. 2001). Our results show that inbreedingdepression is regularly occurring in naturalpopulations following establishment by only onemated female, especially if the population remainssmall for several generations. The combinationof lowered egg-hatching rate and low larvalsurvival amplified by elevated overwintermortality is very likely to be the mechanism leadingto population extinction in inbred populations inthe wild. These results underline the importanceof interactions between genetic and ecologicalfactors in determining population fitness andextinction risk. Interestingly, most of the long-term studies of inbreeding depression in the wildreport a similar interaction between unfavourableecological conditions and the strength of

13

Page 14: Behavioural, population, and genetic processes affecting

inbreeding depression (Coltman et al. 1999; Kelleret al. 1994).

Life history evolution in metapopulations

The number of studies on life history evolutionin the metapopulation context is surprisingly small(Ronce and Olivieri 2004). The field is largelydominated by theory, with very little empiricalwork, and it is dealing almost exclusively withquestions related to the evolution of dispersal(Ronce and Olivieri 2004). The two last chaptersof my thesis deal with the evolution of life historytraits. The data come from a large outdoor cageexperiment that was conducted in summer 2003,in which matings and ovipositions were recordedintensively, and the offspring were reared untildiapause (Hanski et al. 2006). Such data areexceptionally well suited to the study of naturalselection, because it allows determining thereproductive success of all individuals. The fourthchapter of my thesis deals with the evolution ofmultiple mating in females. Although multiplemating has been assumed to be rare in theGlanville fritillary butterfly, a large proportion offemales were found to mate multiply, with up tothree males, in the population cage. I usedmicrosatellite markers to determine the spermprecedence pattern, and I found a very stronglast-male sperm precedence. The data was usedto study multiple matings in females, and also todetermine the reproductive success of males inorder to study protandry.

Both protandry (the earlier emergence of malesthan females) and polyandry (multiple mating infemales) are common life-history phenomenon ininsects (Morbey and Ydenberg 2001; Thornhilland Alcock 1983). Protandry was first describedby Darwin in 1871, who postulated that protandrycould have evolved by natural selection formaximizing the reproductive success of males(Darwin 1871). In addition to adaptiveexplanations of protandry, alternative incidentalhypotheses have been postulated, wherebyprotandry would arise as a by-product ofselection acting on females, but not on males, forincreased body size. No selection would beexpected on the timing of male emergence, andmales emerging at different times should not differ

in mating success (Baughman 1991). In addition,under this hypothesis, there should be a trade-off between developmental time and adult size(Nylin et al. 1993). These questions have mainlybeen addressed with theoretical models, andempirical data to support any of these hypothesesare scarce and controversial (Del Castillo andNunez-Farfan 1999; Maklakov et al. 2004; Nylinet al. 1993; Wedell 1992), mainly due to thedifficulties involved in estimating mating successin insects. My results demonstrate that incidentalhypotheses to explain protandry are unlikely, asthere was no correlation between adult size anddevelopment time in females or males. Whenestimating the strength of natural selection withthese data, I could show that protandry isadaptive in males, although the strength andshape of natural selection is likely to vary in spaceand time. These results contradict predictions ofevolutionary stable strategy models and supportmore deterministic models developed to explainthe evolution of protandry. Nevertheless, theoptimal level of protandry is likely to be influenceby population density. At very low populationdensities pronounced protandry might bedisadvantageous, as it would decrease thephenological overlap between potential mates.Also the metapopulation structure might affectthe optimal level of protandry. Males that emergelate will be alive later in the breeding season, andwill therefore be more likely to mate with animmigrant female and produce high-qualityoutbred offspring. Although females are expectedto have already mated before dispersal, asignificant proportion of females was found tomate multiply in the outdoor cage experiment. Itwould be interesting to compare levels ofprotandry in metapopulations and in morecontinuous populations at different populationdensities.

The last chapter of my thesis deals with theevolution of multiple mating in females. For males,reproductive success is expected to increaselinearly with the number of mates, but theadvantages of multiple mating for females are lessclear (Yasui 1997). Mating is costly to females,because of time and energy costs, and oftenbecause of increased risk of predation, injury orinfection (Blanckenhorn et al. 2002; Chapman et

14

Page 15: Behavioural, population, and genetic processes affecting

al. 1995). Multiple mating by females has beenexplained in terms of direct benefits, particularlyin species in which males provide females with anutrient-rich ejaculate (Arnqvist and Nilsson2000), and in terms of indirect genetic benefits(Fedorka and Mousseau 2002; Jennions andPetrie 2000; Kozielska et al. 2004; Newcomer et al.1999). Indirect genetic benefits include both theinheritance of good genes by the offspring(Andersson 1994; Johnstone 1995; Wedell andTregenza 1999), the avoidance of geneticincompatibility (Foerster et al. 2003; Zeh and Zeh1996), and the increase in offspring geneticdiversity (Watson 1991; Yasui 1998). In manyspecies of Lepidoptera, the ejaculate transferredby the male to the female during mating functionsas a nuptial gift which increases fecundity, eggsize or longevity in multiply mated females (Boggs1990; Wiklund et al. 1993). In a review of 122experimental insect studies, (Arnqvist andNilsson 2000) found that polyandry could beexplained by direct benefits alone, even in specieswithout nuptial gift. On the other hand, in theirreview of polyandry and fecundity inLepidoptera, (Torres-Vila et al. 2004) found thatre-mating had no detectable effects on fecundityin descriptive studies of monandrous species.Similarly, there is no evidence that polyandryleads to direct benefits in Drosophilamelanogaster (Brown et al. 2004).

When pre-copulatory cues allow females toidentify the genetic quality of potential mates,they can potentially engage in additional matingswith genetically superior (e.g. (Hasselquist et al.1996; Kempenaers et al. 1997; Pitcher et al. 2003)or more compatible males (Garner and Schmidt2003; Masters et al. 2003). When females cannotreliably identify the genetic quality of males, theymay rely on post-copulatory cues to increase thebias in fertilization by using sperm that will confertheir offspring with the highest genetic benefits(Jennions and Petrie 2000). Males of highergenetic quality are more likely to produce high-quality ejaculate and have a higher share ofpaternity in sperm competition (Arnqvist 1989;Parker 1990). Post-copulatory paternity biasingfor greater compatibility has also been observed(Clark et al. 1999; Evans and Marshall 2005), forexample through preferential destruction of

genetically more similar sperm (Bishop 1996). Inaddition, there is evidence that post-copulatorypaternity biasing increases offspring viability(Konior et al. 2001). As a result of post-copulatoryinbreeding avoidance, offspring survivalincreased with the number of mates both in theEuropean adder (Madsen et al. 1992) and in theEuropean sand lizard (Olsson et al. 1994). In alaboratory experiment with the yellow dung fly,superior sperm competitors sired higher-qualityoffspring (Hosken and Stockley 2003).

Finally, genetic bet-hedging (Gillespie 1973;Gillespie 1974; Gillespie 1975; Gillespie 1977;Hopper 1999; Seger and Brockman 1987) isanother mechanism that could explain polyandry,especially when females mate indiscriminately(Fox and Rauter 2003; Yasui 1998; Yasui 2001).Bet-hedging is expected to reduce variance infitness among individuals within one generationand to increase the geometric mean fitness ofpolyandrous females relative to that ofmonandrous females. By mating with severalmales, females can avoid having all their offspringfathered by a low-quality or an incompatible male(Fox and Rauter 2003; Jennions and Petrie 2000;Yasui 1998; Yasui 2001).

One reason why bet-hedging can be important isthe metapopulation structure of the Glanvillefritillary in the Åland Islands (Hanski 1999;Nieminen et al. 2004). When local populationscan be so small that a single mated female canestablish an entire new local population (Hanski1999; Hanski et al. 1995), ensuring that at leastsome offspring are viable through variancereduction is a valid argument in favour of bet-hedging (Yasui 1998). This can be achieved eitherby mating multiply before dispersal, oralternatively mating in the new patch if males areencountered there. Re-mating after dispersalwould provide the added benefit of mating withan unrelated male, which is likely to increaseoffspring fitness. The benefits of bet-hedgingcould be accentuated by the fact that the groupsof gregarious larvae have to exceed a thresholdsize before they are likely to survive throughoutthe larval stage (Kuussaari 1998), which makes itparticularly important to bet-hedge in order to

15

Page 16: Behavioural, population, and genetic processes affecting

ensure that at least one larval group will be largeenough to survive the winter.

In the last chapter of my thesis, in a study I didtogether with Hanna Kokko, we derivepredictions of the bet-hedging hypothesis for theGlanville fritillary. The egg-laying and survivalpatterns are in line with the predictions we derive,supporting the hypothesis that multiple matingin the Glanville fritillary presents a rare case ofwithin-generation bet-hedging. Because thisstudy was based on the data from the outdoorcage experiment in which butterflies wereunrelated, any incompatibility effects due toinbreeding depression are likely to beunderestimated. Therefore, both the egg-layingpattern of polyandrous females and the resultingreduced variance in their life-time reproductivesuccess suggest that within-generation bet-hedging is an unusually likely candidate forexplaining the occurrence of polyandry as asuccessful evolutionary strategy in ametapopulation context.

Conclusions

The results of my thesis bring new understandingto the consequences of extinction-colonizationdynamics on genetic and evolutionary processesin metapopulations. I focused primarily oncolonizations (II and III), because in the Glanvillefritillary metapopulation in the Åland Islandscolonizations (and local extinctions) are veryfrequent and the system presents a prime exampleof colonizations by only few females (II). Theresults show that population establishment byonly one mated female may lead to rapidinbreeding depression, which combined withelevated overwinter mortality due to reduced sizeof larval groups may ultimately lead to populationextinction (III, Figure 1). These adverse effects oflow genetic diversity are amplified even furtherby the fact that females usually mate in their natalpopulation before dispersal (II, Figure 1). On theother hand, population establishment by severalfemales is expected to increase greatly the geneticdiversity of the newly-established populations,especially as the founder females are usuallyunrelated (II, Figure 1). In the next generation,

mating among these unrelated groups should leadto high offspring fitness (III, Figure 1). Migrationamong local populations may additionally rescuesmall inbred populations by introducing newgenetic variation (III, Figure 1).

Multiple mating in the Glanville fritillary butterflyseems to be much more common than previouslyassumed, as a large proportion of females wasfound to mate with two or three males in a largeoutdoor experiment (Hanski et al. 2006). Multiplemating in the natal population could alleviate theadverse effects associated with populationestablishment through a bet-hedging mechanism,which would allow the least related male (or thebest male in some other regard) to sire moreoffspring (V). If females disperse to an existingpopulation, re-mating after dispersal wouldharbour the added benefit of increased offspringfitness resulting from low parental relatedness(III). Therefore, multiple mating is likely to bemaintained in the metapopulation context as asuccessful evolutionary strategy (Figure 1), inspite of the costs associated with it. Inbreedingdepression associated with populationestablishment is also likely to have an influenceon the evolution of other important life-historytraits. One obvious trait is the evolution ofdispersal. A metapopulation structure withfrequent colonizations resulting in small inbredpopulations would favour highly dispersiveindividuals, because dispersers into suchpopulations would enjoy strongly increasedreproductive success (III). Other important life-history traits, like protandry, are also likely to beinfluenced by the metapopulation structure (IV,Figure 1). This demonstrates the importance ofintegrating genetic and ecological factors.Inbreeding depression and larval-group sizedependent mortality due to variousenvironmental factors combine to play a key roleboth in determining the extinction risk of smallpopulations and in influencing the evolution ofmultiple mating in females (III and V).

16

Page 17: Behavioural, population, and genetic processes affecting

Population establishmentby only one female

Population establishmentby several, unrelated females

Females mate in their natal populationÞ Low genetic diversity of their offspring

ORMultiple mating in the natal population,or remating after dispersalÞ High genetic diversity of offspring, advantages through bet-hedging

Rescue by immigrationMatings between residents and immigrantsOffspring with high fitnessProtandry advantageous

Several generations ofstrong inbreeding

Low egg hatching rateLow larval survivalHigh overwinter mortalitySmall population sizeProtandry unfavorableHigh extinction risk

Low genetic diversity High genetic diversity

Figure 1. Effects of colonization pattern, female matings and immigration on the maintenance ofgenetic diversity in metapopulations

17

Page 18: Behavioural, population, and genetic processes affecting

Box 1. Parentage analysis methods modified after (Jones and Ardren 2003; Van de Casteele et al. 2001)

Genetic studies of parentage play a major role in the study of evolution and behavioural ecology.There are several methods available for reconstructing patterns of parentage (see below), and theappropriate technique for data analysis is dictated in large part by the type of samples that can beobtained from the study system and the available molecular markers. The best-case scenario is onein which large groups of offspring can be collected from known mated pairs (e.g. verifying parentoffspring relationships to determine sperm precedence pattern in multiply mated females, IV andV). The situation becomes more difficult as the completeness of the sample diminishes (e.g.reconstruction of parental genotypes from groups of offspring, II).

Exclusion based on Mendelian rules of inheritance uses incompatibilities between parents and offspring to reject candidate parents best method if all or some of the potential parents are sampled and if no large sib- groups areavailable needs highly polymorphic markers and is very sensitive to mutations and genotyping errors

Categorical allocation and fractional allocation uses likelihood-based approaches to select the most likely parents from a pool of non-excludedcandidate parents (Meagher and Thompson 1986) fractional allocation assigns some fraction, between 0 and 1, to all non-excluded parents, andlikelihoods are calculated in the same way as in the categorical allocation method (Devlin et al.1988) can be used in addition to exclusion methods if the markers are not polymorphic enough andthere are still several non-excluded parents remaining allows for some amount of mutations and genotyping errors

Parental reconstruction uses multilocus genotypes of parents and offspring to reconstruct the genotype of unknownparents (Jones 2001) needs sampling of large groups of full- or half- sibs can be used even if candidate parents are not known needs highly polymorphic markers

Relatedness estimation neither of the parents is known, candidate parents cannot be sampled and no large sib-groupsavailable, parentage cannot be reconstructed, relatedness techniques must be used several relatedness estimators have been proposed: a similarity index (Li et al. 1993), a regressionbased estimator (Queller and Goodnight 1989), and a correlation-based moment estimator (Lynchand Ritland 1999) typically large errors of inference (Lynch and Ritland 1999; Ritland 1996)

18

Page 19: Behavioural, population, and genetic processes affecting

Acknowledgements

I would like to express my deepest gratitude tomy supervisor, Ilkka Hanski, for providing me allthe support and guidance I needed while givingme the freedom to find my own path, and for notlosing faith in what felt at times like an impossibletask. This thesis would never have beencompleted without his full commitment andsupport through difficult times. His enthusiasm,genuine interest in science, and outstandingexpertise were a source of inspiration on my wayinto becoming an independent scientist. I havebeen given the rare opportunity to work andinteract with exceptional scientists, and I wouldparticularly like to thank Ilkka Hanski, HannaKokko and Otso Ovaskainen for the fruitfulcollaborations. Working with them was veryedifying and a real pleasure. There are manypeople in the department I would like to thank forstimulating scientific conversations and usefulcomments on my manuscripts, particularly OscarGaggiotti, Christoph Haag, Tomas Roslin andJuha Merilä.

I am deeply grateful to Marjo Saastamoinen, whoby conducting the cage experiment in the ÅlandIslands provided me the data for the last twochapters, and made thereby this thesis possible.I would also like to thank Marjo for her friendship.Many thanks go to Marko Nieminen, who hasbeen conducting for many years the annualsurveys on which this thesis is building up, andto Evgeniy Meyke, for taking care of the hugedatabase and always being ready to producecountless maps and tables. I would also like tothank Anu Väisänen, Tapio Gustafsson and NinaBlomqvist for helping me with so many practicalissues. My warmest thanks go to Mimma Ekbladfor assisting me whenever I had to fight with thecomputer (while having to bear my obnoxiousmood ), for editing this thesis and designingthe cover, and finally for a friendship that will lastbeyond the completion of this thesis.

I would like to thank all the members of themetapopulation research group for providingsuch a pleasant and stimulating workenvironment, and for sharing time with me outsideof work. I consider so many my friends, that it

would be impossible to mention all of them. Iwould nevertheless like to thank Tarja for manyyears of friendship and for the moral supportduring the time just before my defence. Manythanks go to Céline, for becoming such a goodfriend in such a short time, as well as Heidi, Ace,Luisa, Itsuro, Andrès, Mike, Varpu and so manyothers for all the parties, the fun, and for theirfriendship. Additional thanks go to my other officemates, Jenni and Sofia, for their friendship andsupport. I would like to thank all my friends outsideof work, particularly Claudia for luring me toFinland and for over ten years of friendship, aswell as Anna and Carole. Many thanks to Katrinand Erkki for letting me ride their beautiful horses.Special thanks go to my parents, my sister Maysaand my brother Joe, for always supporting meand believing in me, across oceans andcontinents. Special thanks also go to a very specialfriend, Patrik, for bringing music into my life, forthis amazing summer and every moment we share.

References

Allendorf, F. W. 1986. Genetic Drift and the Lossof Alleles Versus Heterozygosity. ZooBiology 5:181-190.

Allendorf, F. W., and R. F. Leary. 1986.Heterozygosity and fitness in naturalpopulations of animals. in M. E. Soulé, ed.Conservation Biology, The Science ofScarcity and Diversity. Sinauer AssociatesInc. , Sunderland, Massachusetts.

Amos, W., J. W. Wilmer, K. Fullard, T. M. Burg, J.P. Croxall, D. Bloch, and T. Coulson. 2001.The influence of parental relatedness onreproductive success. Proceedings of theRoyal Society of London Series B-Biological Sciences 268:2021-2027.

Andersson, M. 1994. Sexual selection. PrincetonUniversity Press, Princeton.

Angers, B., A. Estoup, and P. Jarne. 2000.Microsatellite size homoplasy, SSCP, andpopulation structure: A case study in thefreshwater snail Bulinus truncatus.Molecular Biology and Evolution 17:1926-1932.

19

Page 20: Behavioural, population, and genetic processes affecting

Arnqvist, G. 1989. Multiple Mating in a WaterStrider - Mutual Benefits or IntersexualConflict. Animal Behaviour 38:749-756.

Arnqvist, G., and T. Nilsson. 2000. The evolutionof polyandry: multiple mating and femalefitness in insects. Animal Behaviour 60:145-164.

Baughman, J. F. 1991. Do Protandrous MalesHave Increased Mating Success - the Caseof Euphydryas-Editha. American Naturalist138:536-542.

Bishop, J. D. D. 1996. Female control of paternityin the internally fertilizing compoundascidian Diplosoma listerianum .1.Autoradiographic investigation of spermmovements in the female reproductivetract. Proceedings of the Royal Society ofLondon Series B-Biological Sciences263:369-376.

Blanckenhorn, W. U., D. J. Hosken, O. Y. Martin,C. Reim, Y. Teuschl, and P. I. Ward. 2002.The costs of copulating in the dung flySepsis cynipsea. Behavioral Ecology13:353-358.

Boggs, C. L. 1990. A General-Model of the Roleof Male-Donated Nutrients in FemaleInsects Reproduction. American Naturalist136:598-617.

Boggs, C. L., and M. Nieminen. 2004. Checkerspotreproductive biology. . Pp. 92-111 in I.Hanski and P. R. Ehrlich, eds. On the wingsof checkerspots. A model system forpopulation biology. Oxford Univ. Press,Oxford.

Brown, W. D., A. Bjork, K. Schneider, and S.Pitnick. 2004. No evidence that polyandrybenefits females in Drosophilamelanogaster. Evolution 58:1242-1250.

Buchan, J. C., E. A. Archie, R. C. Van Horn, C. J.Moss, and S. C. Alberts. 2005. Locuseffects and sources of error in noninvasivegenotyping. Molecular Ecology Notes5:680-683.

Callen, D. F., A. D. Thompson, Y. Shen, H. A.Phillips, R. I. Richards, J. C. Mulley, and G.R. Sutherland. 1993. Incidence and Originof Null Alleles in the (Ac)N MicrosatelliteMarkers. American Journal of HumanGenetics 52:922-927.

Chapman, T., L. F. Liddle, J. M. Kalb, M. F. Wolfner,and L. Partridge. 1995. Cost of Mating inDrosophila-Melanogaster Females IsMediated by Male Accessory-GlandProducts. Nature 373:241-244.

Charlesworth, D., and B. Charlesworth. 1987.Inbreeding Depression and ItsEvolutionary Consequences. AnnualReview of Ecology and Systematics 18:237-268.

Clark, A. G., D. J. Begun, and T. Prout. 1999.Female x male interactions in Drosophilasperm competition. Science 283:217-220.

Coltman, D. W., J. G. Pilkington, J. A. Smith, and J.M. Pemberton. 1999. Parasite-mediatedselection against inbred Soay sheep in afree-living, island population. Evolution53:1259-1267.

Coulson, T. N., J. M. Pemberton, S. D. Albon, M.Beaumont, T. C. Marshall, J. Slate, F. E.Guinness, and T. H. Clutton-Brock. 1998.Microsatellites reveal heterosis in red deer.Proceedings of the Royal Society ofLondon Series B-Biological Sciences265:489-495.

Crnokrak, P., and D. A. Roff. 1999. Inbreedingdepression in the wild. Heredity 83:260-270.

Crow, J. F. 1993. Mutation, mean fitness andgenetic load. Oxf. Surve. Evol. Biol. 9:3-42.

Cruz, F., M. Perez, and P. Presa. 2005. Distributionand abundance of microsatellites in thegenome of bivalves. Gene 346:241-247.

Darwin, C. 1871. The descent of man and selectionin relation to sex. J. Murray, London.

Dawson, M. N., K. A. Raskoff, and D. K. Jacobs.1998. Field preservation of marineinvertebrate tissue for DNA analyses.Molecular Marine Biology andBiotechnology 7:145-152.

Del Castillo, R. C., and J. Nunez-Farfan. 1999.Sexual selection on maturation time andbody size in Sphenarium purpurascens(Orthoptera : Pyrgomorphidae): Correlatedresponse to selection. Evolution 53:209-215.

Delmotte, F., N. Leterme, and J. C. Simon. 2001.Microsatellite allele sizing: Differencebetween automated capillaryelectrophoresis and manual technique.Biotechniques 31:810-+.

20

Page 21: Behavioural, population, and genetic processes affecting

Devlin, B., K. Roeder, and N. C. Ellstrand. 1988.Fractional Paternity Assignment -Theoretical Development and Comparisonto Other Methods. Theoretical and AppliedGenetics 76:369-380.

Ebert, D., C. Haag, M. Kirkpatrick, M. Riek, J. W.Hottinger, and V. I. Pajunen. 2002. Aselective advantage to immigrant genes ina Daphnia metapopulation. Science295:485-488.

Ehrlich, P. R., and I. Hanski. 2004. On the Wingsof Checkerspots. Oxford University Press,Oxford, UK.

Eisen, J. A. 1999. Mechanistic basis formicrosatelllite instability. Pp. 34-48 in D. B.S. Goldstein, C., ed. Microsatellites:Evolution and Applications. OxfordUniversity Press, Oxford, UK.

Ellegren, H. 2004. Microsatellites: Simplesequences with complex evolution. NatureReviews Genetics 5:435-445.

Estoup, A., and J. M. Cornuet. 1999. Microsatelliteevollution: inferences from population data.Pp. 49-64 in D. B. S. Goldstein, C., ed.Microsatellites: Evolution and Applications.Oxford University Press, Oxford, UK.

Evans, J. P., and D. J. Marshall. 2005. Male-by-female interactions influence fertilizationsuccess and mediate the benefits ofpolyandry in the sea urchin Heliocidariserythrogramma. Evolution 59:106-112.

Falconer, D. S., and T. F. C. Mackay. 1996.Introduction to Quantitative Genetics.,Longman, Harlow, England.

Fedorka, K. M., and T. A. Mousseau. 2002.Material and genetic benefits of femalemultiple mating and polyandry. AnimalBehaviour 64:361-367.

Flanagan, N. S., M. J. Blum, A. Davison, M.Alamo, R. Albarran, K. Faulhaber, E.Peterson, and W. O. McMillan. 2002.Characterization of microsatellite loci inneotropical Heliconius butterflies.Molecular Ecology Notes 2:398-401.

Foerster, K., K. Delhey, A. Johnsen, J. T. Lifjeld,and B. Kempenaers. 2003. Females increaseoffspring heterozygosity and fitnessthrough extra-pair matings. Nature 425:714-717.

Fox, C. W., and C. M. Rauter. 2003. Bet-hedgingand the evolution of multiple mating.Evolutionary Ecology Research 5:273-286.

Frankham, R. 2005. Genetics and extinction.Biological Conservation 126:131-140.

Gaggiotti, O. E. 2003. Genetic threats topopulation persistence. Annales ZoologiciFennici 40:155-168.

Gaggiotti, O. E., O. Lange, K. Rassmann, and C.Gliddon. 1999. A comparison of twoindirect methods for estimating averagelevels of gene flow using microsatellite data.Molecular Ecology 8:1513-1520.

Garner, T. W. J., and B. R. Schmidt. 2003.Relatedness, body size and paternity in thealpine newt, Triturus alpestris. Proceedingsof the Royal Society of London Series B-Biological Sciences 270:619-624.

Gillespie. 1973. Natural-Selection with VaryingSelection Coefficients - Haploid Model.Genetical Research 21:115-120.

Gillespie. 1974. Natural-Selection for within-Generation Variance in Offspring Number.Genetics 76:601-606.

Gillespie, J. H. 1975. Natural-Selection for within-Generation Variance in Offspring Number2. Discrete Haploid Models. Genetics81:403-413.

Gillespie, J. H. 1977. Natural-Selection forVariances in Offspring Numbers - NewEvolutionary Principle. American Naturalist111:1010-1014.

Glemin, S., J. Ronfort, and T. Bataillon. 2003.Patterns of inbreeding depression andarchitecture of the load in subdividedpopulations. Genetics 165:2193-+.

Goodnight, C. J. 2004. Metapopulationquantitative genetics: the quantitativegenetics of population differentiation in I.Hanski and O. E. Gaggiotti, eds. Ecology,Genetics, and Evolution ofMetapopulations. Elsevier Academic Press.

Haag, C. R., J. W. Hottinger, M. Riek, and D. Ebert.2002. Strong inbreeding depression in aDaphnia metapopulation. Evolution 56:518-526.

Haikola, S., W. Fortelius, R. B. O´Hara, M.Kuussaari, N. Wahlberg, I. J. Saccheri, M.C. Singer, and I. Hanski. 2001. Inbreeding

21

Page 22: Behavioural, population, and genetic processes affecting

depression and the maintenance of geneticload in Melitea cinxia metapopulations.Conservation Genetics 2:325-335.

Hanski, I. 1999. Metapopulation ecology. OxfordUniversity Press, Oxford.

Hanski, I., C. J. Breuker, K. Schops, R. Setchfield,and M. Nieminen. 2002. Population historyand life history influence the migration rateof female Glanville fritillary butterflies.Oikos 98:87-97.

Hanski, I., and M. Heino. 2003. Metapopulation-level adaptation of insect host plantpreference and extinction-colonizationdynamics in heterogeneous landscapes.Theoretical Population Biology 64:281-290.

Hanski, I., T. Pakkala, M. Kuussaari, and G. C. Lei.1995. Metapopulation Persistence of anEndangered Butterfly in a FragmentedLandscape. Oikos 72:21-28.

Hanski, I., M. Saastamoinen, and O. Ovaskainen.2006. Dispersal related life-history trade-offs in a butterfly metapopulation. Journalof animal Ecology 75:91-100.

Harper, G. L., S. Piyapattanakorn, D. Goulson, andN. Maclean. 2000. Isolation ofmicrosatellite markers from the Adonis bluebutterfly (Lysandra bellargus). MolecularEcology 9:1948-1949.

Hasselquist, D., S. Bensch, and T. vonSchantz.1996. Correlation between male songrepertoire, extra-pair paternity and offspringsurvival in the great reed warbler. Nature381:229-232.

Hedgecock, D., G. Li, S. Hubert, K. Bucklin, and V.Ribes. 2004. Widespread null alleles andpoor cross-species amplification ofmicrosatellite DNA loci cloned from thePacific oyster, Crassostrea gigas. Journalof Shellfish Research 23:379-385.

Hedrick, P. W., and S. T. Kalinowski. 2000.Inbreeding depression in conservationbiology. Annual Review of Ecology andSystematics 31:139-162.

Hedrick, P. W., and P. S. Miller. 1992. ConservationGenetics - Techniques and Fundamentals.Ecological Applications 2:30-46.

Higgins, K., and M. Lynch. 2001. Metapopulationextinction caused by mutationaccumulation. Proceedings of the National

Academy of Sciences of the United Statesof America 98:2928-2933.

Hoffman, J. I., and W. Amos. 2005. Microsatellitegenotyping errors: detection approachesfcommon sources and consequences forpaternal exclusion. Molecular Ecology14:599-612.

Hopper, K. R. 1999. Risk-spreading and bet-hedging in insect population biology.Annual Review of Entomology 44:535-560.

Hosken, D. J., and P. Stockley. 2003. Benefits ofpolyandry: A life history perspective. Pp.173-194. Evolutionary Biology, Vol 33.

Jennions, M. D., and M. Petrie. 2000. Why dofemales mate multiply? A review of thegenetic benefits. Biological Reviews 75:21-64.

Johnstone, R. A. 1995. Sexual Selection, HonestAdvertisement and the Handicap Principle- Reviewing the Evidence. BiologicalReviews of the Cambridge PhilosophicalSociety 70:1-65.

Jones, A. G. 2001. GERUD1.0: a computer programfor the reconstruction of parentalgenotypes from progeny arrays usingmultilocus DNA data. Molecular EcologyNotes 1:215-218.

Jones, A. G., and W. R. Ardren. 2003. Methods ofparentage analysis in natural populations.Molecular Ecology 12:2511-2523.

Keller, L. F., P. Arcese, J. N. M. Smith, W. M.Hochachka, and S. C. Stearns. 1994.Selection against Inbred Song SparrowsDuring a Natural-Population Bottleneck.Nature 372:356-357.

Keller, L. F., and D. M. Waller. 2002. Inbreedingeffects in wild populations. Trends inEcology & Evolution 17:230-241.

Kempenaers, B., G. R. Verheyren, and A. A.Dhondt. 1997. Extrapair paternity in theblue tit (Parus caeruleus): female choice,male characteristics, and offspring quality.Behavioral Ecology 8:481-492.

Keyghobadi, N., J. Roland, and C. Strobeck. 1999.Influence of landscape on the populationgenetic structure of the alpine butterflyParnassius smintheus (Papilionidae).Molecular Ecology 8:1481-1495.

22

Page 23: Behavioural, population, and genetic processes affecting

Kirkpatrick, M., and N. H. Barton. 1997. Evolutionof a species range. American Naturalist150:1-23.

Konior, M., J. Radwan, and M. Kolodziejczyk.2001. Polyandry increases offspringfecundity in the bulb mite. Evolution55:1893-1896.

Kozielska, M., A. Krzeminska, and J. Radwan.2004. Good genes and the maternal effectsof polyandry on offspring reproductivesuccess in the bulb mite. Proceedings ofthe Royal Society of London Series B-Biological Sciences 271:165-170.

Kuussaari, M. 1998. Biology of the Glanvillefritillary butterfly (Melitaea cinxia). PhDthesis, Department of Ecology andSystematics, University of Helsinki.

Landry, P. A., M. T. Koskinen, and C. R. Primmer.2002. Deriving evolutionary relationshipsamong populations using microsatellitesand (delta mu)(2): All loci are equal, butsome are more equal than others. Genetics161:1339-1347.

Li, C. C., D. E. Weeks, and A. Chakravarti. 1993.Similarity of DNA Fingerprints Due toChance and Relatedness. Human Heredity43:45-52.

Li, Y. C., A. B. Korol, T. Fahima, A. Beiles, and E.Nevo. 2002. Microsatellites: genomicdistribution, putative functions andmutational mechanisms: a review. MolecularEcology 11:2453-2465.

Lynch, M., and K. Ritland. 1999. Estimation ofpairwise relatedness with molecularmarkers. Genetics 152:1753-1766.

Madsen, T., R. Shine, J. Loman, and T. Hakansson.1992. Why Do Female Adders Copulate SoFrequently. Nature 355:440-441.

Maklakov, A. A., T. Bilde, and Y. Lubin. 2004.Sexual selection for increased male bodysize and protandry in a spider. AnimalBehaviour 68:1041-1048.

Maruyama, T. 1970. On Fixation Probability ofMutant Genes in a Subdivided Population.Genetical Research 15:221-&.

Masters, B. S., B. G. Hicks, L. S. Johnson, and L.A. Erb. 2003. Genotype and extra-pairpaternity in the house wren: a rare-male

effect? Proceedings of the Royal Societyof London Series B-Biological Sciences270:1393-1397.

Mayr, E. 1963. Animal Species and Evolution.Belknap Press, Harvard.

Meagher, T. R., and E. Thompson. 1986. TheRelationship between Single Parent andParent Pair Genetic Likelihoods inGenealogy Reconstruction. TheoreticalPopulation Biology 29:87-106.

Meglecz, E., F. Petenian, E. Danchin, A. C. D Acier,J. Y. Rasplus, and E. Faure. 2004. Highsimilarity between flanking regions ofdifferent microsatellites detected withineach of two species of Lepidoptera:Parnassius apollo and Euphydryas aurinia.Molecular Ecology 13:1693-1700.

Meglecz, E., and M. Solignac. 1998. Microsatelliteloci for Parnassius mnemosyne(Lepidoptera). Hereditas 128:179-180.

Mitton, J. B. 1993. Theory and data pertinent tothe relationship between heterozygosityand fitness. Pp. 17-41 in N. W. Thornhill,ed. The natural history of inbreeding andoutbreeding. The University of ChicagoPress.

Morbey, Y. E., and R. C. Ydenberg. 2001.Protandrous arrival timing to breedingareas: a review. Ecology Letters 4:663-673.

Nei, M., T. Maruyama, and R. Chakraborty. 1975.Bottleneck Effect and Genetic-Variability inPopulations. Evolution 29:1-10.

Neigel, J. E. 1997. A comparison of alternativestrategies for estimating gene flow fromgenetic markers. Annual Review of Ecologyand Systematics 28:105-128.

Newcomer, S. D., J. A. Zeh, and D. W. Zeh. 1999.Genetic benefits enhance the reproductivesuccess of polyandrous females.Proceedings of the National Academy ofSciences of the United States of America96:10236-10241.

Nielsen, E. E., M. M. Hansen, D. E. Ruzzante, D.Meldrup, and P. Gronkjaer. 2003. Evidenceof a hybrid-zone in Atlantic cod (Gadusmorhua) in the Baltic and the Danish BeltSea revealed by individual admixtureanalysis. Molecular Ecology 12:1497-1508.

Nieminen, M., M. Siljander, and I. Hanski. 2004.Structure and dynamics of Melitaea cinxia

23

Page 24: Behavioural, population, and genetic processes affecting

metapopulations. Pp. 63-91 in P. R. Ehrlichand I. Hanski, eds. On the wings ofcheckerspots: a model system forpopulation biology. . Oxford UniversityPress, Oxford.

Nieminen, M., M. C. Singer, W. Fortelius, K.Schops, and I. Hanski. 2001. Experimentalconfirmation that inbreeding depressionincreases extinction risk in butterflypopulations. American Naturalist 157:237-244.

Nylin, S., C. Wiklund, P. O. Wickman, and E.Garciabarros. 1993. Absence of Trade-Offsbetween Sexual Size Dimorphism and EarlyMale Emergence in a Butterfly. Ecology74:1414-1427.

Olsson, M., T. Madsen, R. Shine, A. Gullberg,and H. Tegelstrom. 1994. Rewards ofPromiscuity - Reply. Nature 372:230-230.

Paetkau, D., and C. Strobeck. 1995. TheMolecular-Basis and Evolutionary Historyof a Microsatellite Null Allele in Bears.Molecular Ecology 4:519-520.

Pannell, J. R., and B. Charlesworth. 1999. Neutralgenetic diversity in a metapopulation withrecurrent local extinction andrecolonization. Evolution 53:664-676.

Parker, G. A. 1990. Sperm Competition Games -Sneaks and Extra-Pair Copulations.Proceedings of the Royal Society of LondonSeries B-Biological Sciences 242:127-133.

Pemberton, J. 2004. Measuring inbreedingdepression in the wild: the old ways are thebest. Trends in Ecology & Evolution19:613-615.

Pitcher, T. E., B. D. Neff, F. H. Rodd, and L. Rowe.2003. Multiple mating and sequential matechoice in guppies: females trade up.Proceedings of the Royal Society of LondonSeries B-Biological Sciences 270:1623-1629.

Primmer, C. R., T. Raudsepp, B. P. Chowdhary, A.R. Moller, and H. Ellegren. 1997. Lowfrequency of microsatellites in the aviangenome. Genome Research 7:471-482.

Queller, D. C., and K. F. Goodnight. 1989.Estimating Relatedness Using Genetic-Markers. Evolution 43:258-275.

Queller, D. C., J. E. Strassmann, and C. R. Hughes.1993. Microsatellites and Kinship. Trendsin Ecology & Evolution 8:285-&.

Richards, C. M. 2000. Inbreeding depressionand genetic rescue in a plantmetapopulation. American Naturalist155:383-394.

Ritland, K. 1996. Estimators for pairwiserelatedness and individual inbreedingcoefficients. Genetical Research 67:175-185.

Ronce, O., and I. Olivieri. 2004. Life historyevolution in metapopulations in I. Hanskiand O. E. Gaggiotti, eds. Ecology, Genetics,and Evolution of Metapopulations.Elsevier Academic Press.

Saccheri, I., M. Kuussaari, M. Kankare, P. Vikman,W. Fortelius, and I. Hanski. 1998.Inbreeding and extinction in a butterflymetapopulation. Nature 392:491-494.

Schlötterer, C. 2000. Evolutionary dynamics ofmicrosatellite DNA. Chromosoma 109:365-371.

Schlötterer, C. 2004. The evolution of molecularmarkers - just a matter of fashion? NatureReviews Genetics 5:63-69.

Seger, J., and H. J. Brockman. 1987. What is bet-hedging? Oxford Univ. Press

Selkoe, K. A., and R. J. Toonen. 2006.Microsatellites for ecologists: a practicalguide to using and evaluating microsatellitemarkers. Ecology Letters 9:615-629.

Slate, J., P. David, K. G. Dodds, B. A. Veenvliet, B.C. Glass, T. E. Broad, and J. C. McEwan.2004. Understanding the relationshipbetween the inbreeding coefficient andmultilocus heterozygosity: theoreticalexpectations and empirical data. Heredity93:255-265.

Slatkin, M. 1977. Gene Flow and Genetic Drift ina Species Subject to Frequent LocalExtinctions. Theoretical Population Biology12:253-262.

Slatkin, M. 1995. A Measure of PopulationSubdivision Based on Microsatellite AlleleFrequencies. Genetics 139:457-462.

Thornhill, R., and J. Alcock. 1983. The evolutionof insects mating systems. Harvard Univ.Press, Cambridge, MA.

Torres-Vila, L. M., M. C. Rodriguez-Molina, andM. D. Jennions. 2004. Polyandry andfecundity in the Lepidoptera: canmethodological and conceptual

24

Page 25: Behavioural, population, and genetic processes affecting

approaches bias outcomes? BehavioralEcology and Sociobiology 55:315-324.

Wade, M. J., and D. E. McCauley. 1988. Extinctionand Recolonization - Their Effects on theGenetic Differentiation of Local-Populations. Evolution 42:995-1005.

Wahlberg, N. 1995. One day in the life of abutterfly. A study of the biology of theGlanville fritillary Melitaea cinxia. MScthesis. University of Helsinki.

Van de Casteele, T., P. Galbusera, and E.Matthysen. 2001. A comparison ofmicrosatellite-based pairwise relatednessestimators. Molecular Ecology 10:1539-1549.

Van Oosterhout, C., W. F. Hutchinson, D. P. M.Wills, and P. Shipley. 2004. MICRO-CHECKER: software for identifying andcorrecting genotyping errors inmicrosatellite data. Molecular EcologyNotes 4:535-538.

Wang, J. L., and A. Caballero. 1999. Developmentsin predicting the effective size of subdividedpopulations. Heredity 82:212-226.

Watson, P. J. 1991. Multiple Paternity as GeneticBet-Hedging in Female Sierra Dome Spiders,Linyphia-Litigiosa (Linyphiidae). AnimalBehaviour 41:343-360.

Wattier, R., C. R. Engel, P. Saumitou-Laprade, andM. Valero. 1998. Short allele dominance asa source of heterozygote deficiency atmicrosatellite loci: experimental evidence atthe dinucleotide locus Gv1CT in Gracilariagracilis (Rhodophyta). Molecular Ecology7:1569-1573.

Wedell, N. 1992. Protandry and Mate Assessmentin the Wartbiter Decticus-Verrucivorus(Orthoptera, Tettigoniidae). BehavioralEcology and Sociobiology 31:301-308.

Wedell, N., and T. Tregenza. 1999. Successfulfathers sire successful sons. Evolution53:620-625.

Whitlock, M. 2004. Selection and drift inmetapopulations in I. Hanski and O. E.Gaggiotti, eds. Ecology, Genetics andEvolution of Metapopulations. ElsevierAcademic Press.

25

Whitlock, M. C. 2002. Selection, load andinbreeding depression in a largemetapopullation. Genetics 160:1191-1202.

Whitlock, M. C. 2003. Fixation probability andtime in subdivided populations. Genetics164:767-779.

Whitlock, M. C., and N. H. Barton. 1997. Theeffective size of a subdivided population.Genetics 146:427-441.

Wiklund, C., A. Kaitala, V. Lindfors, and J. Abenius.1993. Polyandry and Its Effect on FemaleReproduction in the Green-Veined WhiteButterfly (Pieris-Napi L). BehavioralEcology and Sociobiology 33:25-33.

Yasui, Y. 1997. A good-sperm model can explainthe evolution of costly multiple mating byfemales. American Naturalist 149:573-584.

Yasui, Y. 1998. The genetic benefits of femalemultiple mating reconsidered. Trends inEcology & Evolution 13:246-250.

Yasui, Y. 2001. Female multiple mating as a geneticbet-hedging strategy when mate choicecriteria are unreliable. Ecological Research16:605-616.

Zeh, J. A., and D. W. Zeh. 1996. The evolution ofpolyandry I: Intragenomic conflict andgenetic incompatibility. Proceedings of theRoyal Society of London Series B-Biological Sciences 263:1711-1717.

Zhang, D. X. 2004. Lepidopteran microsatelliteDNA: redundant but promising. Trends inEcology & Evolution 19:507-509.