14
Rapid contemporary evolution and clonal food web dynamics Laura E. Jones 1 , Lutz Becks 1,† , Stephen P. Ellner 1 , Nelson G. Hairston Jr 1 , Takehito Yoshida 2 and Gregor F. Fussmann 3, * 1 Department of Ecology and Evolutionary Biology, Cornell University, Corson Hall, Ithaca, NY 14853-2701, USA 2 Department of General Systems Studies, University of Tokyo, 3-8-1 Komaba, Meguro, Tokyo 153-8902, Japan 3 Department of Biology, McGill University, 1205 Avenue Docteur-Penfield, Montreal, Quebec H3A 1B1, Canada Character evolution that affects ecological community interactions often occurs contemporaneously with temporal changes in population size, potentially altering the very nature of those dynamics. Such eco-evolutionary processes may be most readily explored in systems with short generations and simple genetics. Asexual and cyclically parthenogenetic organisms such as microalgae, cladocerans and rotifers, which frequently dominate freshwater plankton communities, meet these requirements. Multiple clonal lines can coexist within each species over extended periods, until either fixation occurs or a sexual phase reshuffles the genetic material. When clones differ in traits affecting interspecific interactions, within-species clonal dynamics can have major effects on the population dynamics. We first consider a simple predator–prey system with two prey genotypes, parametrized with data on a well-studied experimental system, and explore how the extent of differences in defence against predation within the prey population determine dynamic stability versus instability of the system. We then explore how increased potential for evolution affects the community dynamics in a more general community model with multiple predator and multiple prey genotypes. These examples illustrate how microevolutionary ‘details’ that enhance or limit the potential for heritable phenotypic change can have significant effects on contemporaneous community-level dynamics and the persistence and coexistence of species. Keywords: predator–prey; eco-evolutionary; genetic diversity; rotifers; alga 1. INTRODUCTION The last three decades have seen an accumulation of studies demonstrating that evolutionary change in ecologically important organismal traits often takes place at the same time and pace as ecological dynamics (‘eco-evolutionary dynamics’; Fussmann et al. 2007; Kinnison & Hairston 2007). Species as diverse as single-celled algae, annual plants, birds, fishes, crus- taceans, insects and sheep are found to undergo rapid contemporary evolutionary changes in traits that adapt them within a few generations to changing or new environments ( Reznick et al. 1997; Thompson 1998; Hairston et al. 1999; Sinervo et al. 2000; Cousyn et al. 2001; Reznick & Ghalambor 2001; Grant & Grant 2002; Heath et al. 2003; Yoshida et al. 2003; Olsen et al. 2004; Pelletier et al. 2007; Swain et al. 2007). Thus, the old assumption that evolutionary change is negligible on the time scale of ecological interactions is now demonstrably incorrect. Ultimately, evolution has the potential to transform how we think about and manage ecological systems, with implications for resource management ( Heath et al. 2003; Swain et al. 2007; Strauss et al. 2008), conservation of biodiversity ( Ferrie `re & Couvert 2004; Kinnison & Hairston 2007), invasive species manage- ment ( Lavergne & Molofsky 2007) and ecosystem responses to environmental change ( Hairston et al. 2005; Strauss et al. 2008). However, virtually all of the ecological theories developed to address these practical issues invoke, tacitly or explicitly, a separation of time scales between ecological and evolutionary dynamics (see Ferrie `re & Couvert 2004). Frequently, that is valid, of course, but the accumulating evidence suggests that important questions may be answered incorrectly if we fail to account for the fact that we are studying and managing a ‘moving target’. Here, we use the compound term ‘rapid contem- porary evolution’ to refer to heritable changes within a population which occur at a rate fast enough to affect interspecific interactions while they are taking place. The descriptor ‘rapid’, by itself, leaves ambiguous the frame over which evolutionary rates are compared (rapid species diversification on a geological time scale can be quite slow compared with the rate of heritable changes within a population). The adjective ‘con- temporary’ ( Hendry & Kinnison 1999) identifies the time scale that interests us, but permits the degree of Phil. Trans. R. Soc. B doi:10.1098/rstb.2009.0004 One contribution of 14 to a Theme Issue ‘Eco-evolutionary dynamics’. * Author for correspondence ([email protected]). Present address: Department of Ecology and Evolutionary Biology, 25 Willcocks Street, Toronto, Ontario M5S 3B2, Canada. RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1–14 ARTICLE IN PRESS 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 1 This journal is q 2009 The Royal Society

Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

Phil. Trans. R. Soc. B

doi:10.1098/rstb.2009.0004

ARTICLE IN PRESS

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

Rapid contemporary evolution and clonalfood web dynamics

72

73

74

75

Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson G. Hairston Jr1,

Takehito Yoshida2 and Gregor F. Fussmann3,*

One con

*Autho† Presen25 Willc

RSTB 20

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

1Department of Ecology and Evolutionary Biology, Cornell University,Corson Hall, Ithaca, NY 14853-2701, USA

2Department of General Systems Studies, University of Tokyo, 3-8-1 Komaba, Meguro,Tokyo 153-8902, Japan

3Department of Biology, McGill University, 1205 Avenue Docteur-Penfield, Montreal,Quebec H3A 1B1, Canada

Character evolution that affects ecological community interactions often occurs contemporaneouslywith temporal changes in population size, potentially altering the very nature of those dynamics. Sucheco-evolutionary processes may be most readily explored in systems with short generations andsimple genetics. Asexual and cyclically parthenogenetic organisms such as microalgae, cladoceransand rotifers, which frequently dominate freshwater plankton communities, meet these requirements.Multiple clonal lines can coexist within each species over extended periods, until either fixationoccurs or a sexual phase reshuffles the genetic material. When clones differ in traits affectinginterspecific interactions, within-species clonal dynamics can have major effects on the populationdynamics. We first consider a simple predator–prey system with two prey genotypes, parametrizedwith data on a well-studied experimental system, and explore how the extent of differences in defenceagainst predation within the prey population determine dynamic stability versus instability of thesystem. We then explore how increased potential for evolution affects the community dynamics in amore general community model with multiple predator and multiple prey genotypes. These examplesillustrate how microevolutionary ‘details’ that enhance or limit the potential for heritable phenotypicchange can have significant effects on contemporaneous community-level dynamics and thepersistence and coexistence of species.

Keywords: predator–prey; eco-evolutionary; genetic diversity; rotifers; alga

100

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

1. INTRODUCTIONThe last three decades have seen an accumulation ofstudies demonstrating that evolutionary change inecologically important organismal traits often takesplace at the same time and pace as ecological dynamics(‘eco-evolutionary dynamics’; Fussmann et al. 2007;Kinnison & Hairston 2007). Species as diverse assingle-celled algae, annual plants, birds, fishes, crus-taceans, insects and sheep are found to undergo rapidcontemporary evolutionary changes in traits that adaptthem within a few generations to changing or newenvironments (Reznick et al. 1997; Thompson 1998;Hairston et al. 1999; Sinervo et al. 2000; Cousyn et al.2001; Reznick & Ghalambor 2001; Grant & Grant2002; Heath et al. 2003; Yoshida et al. 2003; Olsenet al. 2004; Pelletier et al. 2007; Swain et al. 2007).Thus, the old assumption that evolutionary change isnegligible on the time scale of ecological interactionsis now demonstrably incorrect.

Ultimately, evolution has the potential to transformhow we think about and manage ecological systems,

tribution of 14 to a Theme Issue ‘Eco-evolutionary dynamics’.

r for correspondence ([email protected]).t address: Department of Ecology and Evolutionary Biology,ocks Street, Toronto, Ontario M5S 3B2, Canada.

090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1

124

125

126

127

128

1

with implications for resource management (Heathet al. 2003; Swain et al. 2007; Strauss et al. 2008),

conservation of biodiversity (Ferriere & Couvert 2004;

Kinnison & Hairston 2007), invasive species manage-

ment (Lavergne & Molofsky 2007) and ecosystem

responses to environmental change (Hairston et al.2005; Strauss et al. 2008). However, virtually all of theecological theories developed to address these practical

issues invoke, tacitly or explicitly, a separation of time

scales between ecological and evolutionary dynamics

(see Ferriere & Couvert 2004). Frequently, that is

valid, of course, but the accumulating evidence

suggests that important questions may be answeredincorrectly if we fail to account for the fact that we are

studying and managing a ‘moving target’.

Here, we use the compound term ‘rapid contem-

porary evolution’ to refer to heritable changes within a

population which occur at a rate fast enough to affect

interspecific interactions while they are taking place.The descriptor ‘rapid’, by itself, leaves ambiguous the

frame over which evolutionary rates are compared

(rapid species diversification on a geological time scale

can be quite slow compared with the rate of heritable

changes within a population). The adjective ‘con-

temporary’ (Hendry & Kinnison 1999) identifies thetime scale that interests us, but permits the degree of

–14

This journal is q 2009 The Royal Society

Page 2: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

Q1

2 L. E. Jones et al. Rapid evolution in food webs

ARTICLE IN PRESS

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

change to be either negligible or dramatic. Our intent isto explore the effect of evolution that both occurswithin a few generations (the time scale of ecologicaldynamics) and is substantial (ecological interactionstrength is importantly altered), i.e. evolution that isboth contemporary and rapid.

Post & Palkovacs (in press) note that a system of eco-evolutionary feedbacks requires both that the pheno-types must affect the environment in which they liveand the environment must select on the distribution ofgenetically based phenotypes. A major component ofecological research has been the identification andmeasurement of interaction strengths among species,and between species and their chemical and physicalenvironments (e.g. Cain et al. 2008; Molles 2008), andPost & Palkovacs (in press) provide a table of fineexamples. At the same time, there is a growingliterature showing that heritable phenotypic differencesamong individuals have substantial importance for howcommunities and ecosystems function (e.g. Johnsonet al. 2006; Whitham et al. 2006; Urban et al. 2008;Palkovacs et al. in press). Finally, extensive reviews ofthe literature (Hendry & Kinnison 1999; Kinnison &Hendry 2001; Palumbi 2001) leave no doubt thatselection driven by strong ecological interactionsresults in marked adaptive evolutionary change thatfrequently takes place within a few generations.

A challenge, then, is to draw together each of thesecomponents within a single system to determine theimportance of rapid contemporary evolution to eco-logical dynamics. Hairston et al. (2005) proposed anapproach, first conceived by Monica Geber, in whichthe relative importance of temporal changes in a featureof the environment and heritable changes within apopulation are assessed in terms of the magnitudes andtheir contributions to ecological dynamics. Theyapplied it to data on Galapagos finches (Grant &Grant 2002), freshwater copepods (Ellner et al. 1999)and laboratory microcosms (Yoshida et al. 2003).Pelletier et al. (2007) applied a similar method toSoay sheep, and Ezard et al. (in press) have used thisapproach on five populations of ungulates. In eachcase, evolutionary contributions to ecological dynamicswere found to be substantial.

Some of the most striking examples of temporalecological dynamics in nature are found in consumer–resource interactions. Theory indicates and empiricaldata confirm that population abundances frequentlyoscillate substantially as interactions vary in strength,and in some cases sign, over time. Many of the clearestexamples of rapid contemporary evolution have beendocumented in these systems including predatory fishand their prey (Hairston & Walton 1986; Reznick et al.1997; Ellner et al. 1999; Hairston & De Meester2008), herbivorous insects and the plants theyconsume (Johnson et al. 2006, in press) and parasitesand pathogens and their hosts (Dube et al. 2002;Decaestecker et al. 2007; Duffy et al. 2007; Eldredet al. 2008). Owing to the propensity to oscillate,consumer–resource interactions are particularly goodmodel systems for detailed laboratory study of eco-evolutionary dynamics over both long (Lenski &Travisano 1994; Boraas et al. 1998) and short term(Meyer & Kassen 2007). Investigations using simple

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB –

Phil. Trans. R. Soc. B

laboratory microcosms of real consumer and resourcespecies combined with mathematical models thatincorporate observed interactions and predict result-ing dynamics have proven to be particularly fruitful foruncovering some of the potential and diversity of eco-evolutionary dynamics (e.g. Yoshida et al. 2003, 2007;Meyer et al. 2006; Jones & Ellner 2007).

In this paper, we explore how evolutionary dynamicsin traits that influence the strength and nature ofinterspecific interactions affect the temporal dynamicsof species, and the reciprocal effects of populationdynamics on maintenance of the heritable variationthat allows evolution to continue. We begin byreviewing some of our relevant prior experimentaland theoretical work on rapid contemporary evolutionin predator–prey systems, and then use extensionsof our previous models to address the followinggeneral questions:

(i) How does evolution affect the dynamics ofecological communities, and their persistence,robustness and stability?

(ii) What are the dynamic possibilities when a widerange of heritable trait variation is present in asystem, and what are the consequences ofdiminished genetic diversity and the correspond-ingly reduced potential for evolutionary change?

Theory linking evolution and ecology has beenpublished for decades, but theory unfettered by datacan predict (and has predicted) any conceivableecological dynamics, including stability, instability,chaos and criticality (the boundary between twodifferent types of dynamics, e.g. stability versus cycles).To make specific predictions possible, our theory isclosely tied to our experimental work on predator–preychemostats that are laboratory analogues of freshwaterplanktonic food webs.

Natural freshwater plankton systems are oftendominated by obligately asexual or facultatively sexualspecies, so our models assume that trait evolutionoccurs through changes in the frequency of differentclonal lineages. However, models for quantitative traitdynamics in sexual species can be very similar to ourmodels for clone-frequency dynamics, and can even bemathematically identical (depending on assumptionsabout the genetic variance for the quantitative trait; seeAbrams & Matsuda 1997; Abrams 2005). As a result,our conclusions might also be relevant to sexuallyreproducing species. Models for clone-frequencychange are also mathematically equivalent to modelsfor multispecies interactions, but some predictionsabout clonal evolution result from constraints on thedifferences among clones within a species, whichtypically would not be realistic if we were modellingcompletely different species (Jones & Ellner 2007).

2. BACKGROUND: PREDATOR AND PREYIN THE CHEMOSTATIn laboratory microcosmscontaininga rotifer,Brachionuscalyciflorus, consuming an alga, Chlorella vulgaris, weobserved predator–prey cycles with the classic quarter-period phase lag between the peaks in prey and predator

pp. 1–14

Page 3: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

0 10 20 30 40 50

0 10 20 30 40 50 60

1

2

3

4(a) (b)

(c) (d )

(e) ( f )

1

2

3

4

1

2

3

4

prey

den

sity

and

trai

t

0 0

5

10

15

20

0

5

10

15

0

5

10

15

20

20

40

60

80

0

0

2

4

6

8

10

12

20

40

60

80

prey

den

sity

and

trai

t

0 50 100 150 200time (days)

prey

den

sity

and

trai

t

5 10 15 20 25 30 35

0

1.0

2.0

3.0

1.0

2.0

3.0

0

1.0

2.0

3.0pr

ey d

ensi

ty

pred

ator

den

sity

0 10 20 30 40 50 60

prey

den

sity

pred

ator

den

sity

40 50 60 70 80time (days)

prey

den

sity

pred

ator

den

sity

Figure 1. Comparison of theoretical and experimental results for rotifer–algal chemostats with the rotifer B. calyciflorus as thepredator (red circles) and the asexual green alga C. vulgaris as the prey (green circles). (a,b) Classical predator–prey cycles when algaeare monoclonal (one genotype) and so cannot evolve (red solid curve, predator, green dashed curve, prey and purple dotted curve,average prey vulnerability to predation with 1Zno defence and 0Zcomplete invulnerability to attack). Data from Yoshida et al.(2003). (c,d ) Evolutionary cycles when prey are genetically variable. Data from Fussmann et al. (2000). (e, f ) Cryptic cycles whenprey are genetically variable with low cost of defence. Data from Yoshida et al. (2007); spectral analysis confirms the presence ofperiodicity in the predator dynamics (P!0.01). Units: 106 cells mlK1 (algae), females mlK1 (rotifers) and mean palatability (preytrait). Additional experimental replicates for (b,d, f ) can be found in the publications cited as the source for the data.

Rapid evolution in food webs L. E. Jones et al. 3

ARTICLE IN PRESS

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

densities when the algal prey were genetically homo-

geneous. By contrast, with a genetically variable prey

population, cycle period increased three- to sixfold, and

predator and prey cycles were exactly out of phase

(Shertzer et al. 2002; Yoshida et al. 2003; figure 1), a

phenomenon that cannot be explained by conventional

predator–prey models. The algae in our microcosms

vary genetically along a trade-off curve between defence

against rotifer predation and ability to compete for

limiting nutrients (Yoshida et al. 2004). As a result, they

have an evolutionary response to fluctuations in rotifer

density and nutrient availability that qualitatively

changes the predator–prey dynamics.

To understand our experimental results, we for-

mulated and analysed a general model for a predator–

prey system with genetic variability in an asexually

reproducing prey species ( Yoshida et al. 2003, 2007;

Jones & Ellner 2004, 2007). Our results show that the

qualitative properties of evolution-driven cycles in our

experimental system are a general consequence of rapid

prey evolution in a predator–prey food chain, occurring

in a specific but biologically relevant region of

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1

Phil. Trans. R. Soc. B

parameter space. Specifically, genotypes with betterdefence against predation become more common whenpredators are abundant, and less common (due topoorer competitive ability) when predators are rare. Wethen confirmed this theoretical prediction experimen-tally, with close qualitative and quantitative agreementbetween experimental results and a refined versionof our evolutionary model (Yoshida et al. 2003;figure 1a–d ). Finally, we verified the postulated trade-off between defence against predation and ability tocompete for scarce nutrients (Yoshida et al. 2004).

A basic model for our experimental system is givenby equation (2.1) below; this model will be the startingpoint for the new theoretical results that we present inthis paper:

dS

dtZ 1KSK

X2

iZ1

ximS

ki CS;

dxi

dtZ xi

mS

ki CSK

piyG

kb CQK1

� �and

dy

dtZ y

GQ

kb CQK1

� �;

ð2:1Þ

–14

Page 4: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

Q2

0.1 0.2 0.3 0.4 0.5

0

0

0.4

0.8

10

30

(a)

(b)

popu

latio

nde

nsity

EC SS CRC

palatability of defended clone, p1

frac

tion

Figure 2. Bifurcation diagrams for a model with two preyclones having palatabilities p1 and p2Z1, with a linear trade-off between p and competitive ability. (a) The minimum andmaximum of prey (green circles) and predator (red circles)populations, once the populations have settled onto theirattractor (an equilibrium or limit cycle). EC, evolutionarycycles; SS, steady state; CRC, classical consumer–resourcecycles. (b) The minimum and maximum of the fraction of thedefended clone, p1, in the prey population.

4 L. E. Jones et al. Rapid evolution in food webs

ARTICLE IN PRESS

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

where x1, x2, y are the abundances of the two preyclones and the predator; S is the amount of a limitingsubstrate required for algal growth; and QZp1x1Cp2x2

is total prey ‘quality’ as perceived by the predator. Timehas been rescaled so that the dilution rate (the fractionof medium replaced each day) is 1, and state variableshave been rescaled so that the rate of substrate supply is1 and a unit of prey consumption yields one netpredator birth.

The prey types are assumed to be two genotypeswithin a single species, differing in two ways which wehave documented in our experimental systems: their‘palatability’ pi which determines their relative risk ofbeing attacked and consumed by the predator, and theirability to compete for scarce resources, representedby ki. There is a trade-off between defence againstpredation and the ability to compete for nutrients whennutrient concentration is low (Yoshida et al. 2004), sothe genotype with the lower value of p has the highervalue of k. The models fitted to our experimental dataalso include predator age structure (Fussmann et al.2000; Yoshida et al. 2003), mortality within thechemostat and a predator functional response thatbecomes linear (type-I) at low food density, but thesefeatures have no qualitative effect apart from stabilizingthe system at very low dilution rates (Jones & Ellner2007). The model’s main qualitative predictions aboutpotential effects of rapid contemporary prey evolutionon predator–prey cycles are also robust to changes in thefunctional forms of the prey and predator functionalresponses ( Jones & Ellner 2007). However, it isimportant for our results that the model does not allowthe predator to preferentially attack and consume theless defended prey type. Prey genotypes are attacked inproportion to their abundance, and better defencemeans a higher chance of surviving an attack. Theseassumptions have been verified experimentally for ourchemostat system (Meyer et al. 2006).

By mathematical and computational analysis of ageneral version of model (2.1), we have shown (Jones &Ellner 2007) that evolution-driven cycles occur whendefence against predation is effective but cheap: p1 ismuch smaller than p2, but k1 is not much larger thank2. When the cost of defence is extremely low, asurprising phenomenon occurs, which we have called‘cryptic dynamics’: the predator cycles in abundance,but the total prey population remains effectivelyconstant through time (figure 1e). We know that thecost of defence is quite low for some genotypes in ouralgal study species (Meyer et al. 2006), and the crypticdynamics predicted by the model have been observedin our experimental system (Yoshida et al. 2007;figure 1 f ). The constancy of total prey abundanceoccurs in the model due to near exact compensationbetween the defended and undefended genotypes.Each genotype is undergoing large changes in abun-dance, but as the cost of defence is made very low, theoscillations of the two genotypes become almost exactlyout of phase with each other, leading to near constancyof their sum (Jones & Ellner 2007). In addition, thecoefficient of variation in the trait becomes smaller andsmaller relative to the coefficient of variation in thepredators (compare the dotted curves in figure 1eversus figure 1c): only the predator continues to have

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB –

Phil. Trans. R. Soc. B

cycles whose peaks are much higher than the troughs.Since defence is demonstrated to be cheap, or evenfree, in a number of natural and laboratory systems,especially those involving organisms with short gener-ation times (Andersson & Levin 1999; Yoshida et al.2004; Gagneux et al. 2006), cryptic dynamics may bemore common than hitherto anticipated.

3. RESULTS I: DYNAMIC EFFECTS OF PREYGENETIC VARIABILITYThe phenomenon of cryptic dynamics shows that thedynamics of a predator–prey interaction are not justdetermined by the presence or absence of contempor-ary evolution. The details matter, and the nature andextent of genetic variability for traits affecting theinteraction are important details.

Our goal in this paper is to explore theoretically howchanges in the suite of genotypes that are present in theprey and predator populations can affect dynamics atthe population and community levels. We begin, in thissection, by considering genetic variation in prey alone,with prey clones arrayed along a trade-off curvebetween competitive ability and defence againstpredation. Given a trade-off curve, how is thequalitative nature of the population-level dynamicsaffected by the presence or absence of specificgenotypes along the curve, such as the presence orabsence of well-defended types, or the number ofalternative clones filling in the range of variation fromlow to high levels of defence? And how does the shapeof the trade-off curve affect these predictions?

(a) Systems with two extreme prey types

A frequent outcome of competition between twoprey clones in our model is selection for two extremetypes: either the very least and most defended among

pp. 1–14

Page 5: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

Q3

Q4

10 20 30 40 50 60 700

10

20

30

40

0

10

20

30

40

50(a)

(b)

time (days)

popu

latio

ns a

ndm

ean

clum

p si

zepo

pula

tions

and

mea

n cl

ump

size

Figure 3. Chemostat experiments (L. Becks, unpublisheddata) with the rotifer B. calyciflorus as the predator (red solidcurve and filled circles) and the asexual green algaChlamydomonas reinhardii as the prey (green dashed curveand open circles) ((a) weaker defence: SS and (b) strongerdefence: cycles). The circles and triangles are experimentalmeasurements, and the curves are a smooth of the data bylocal polynomial regression. The purple curve (dotted curveand triangles) shows the mean clump size of the algalpopulation. See the text for discussion of clumping as ananti-predator defence trait. Units are individuals mlK1 forpredator population density, 104 cells mlK1 for prey popu-lation density and mean number of cells per clump.

Rapid evolution in food webs L. E. Jones et al. 5

ARTICLE IN PRESS

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

538

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

the clones present in the population, or clones verynear to those extremes. In such cases, the range ofvariation present (the values of p1 and p2) determinesthe outcome, and the shape of the trade-off curveis irrelevant.

Figure 2 summarizes how the population dynamicsare affected by varying the range of palatability whenonly two extreme clones are present. When the better-defended clone has very effective defence ( p1%0.1),theory predicts the type of evolutionary cycles (EC,figure 2) observed by Yoshida et al. (2003), or possiblycryptic cycles (Yoshida et al. 2007), depending on thecost of defence. When defence is less effective, thesystem goes to a steady state (SS, figure 2). For anarrow range of p1-values (e.g. p1z0.14 in figure 2),both prey types coexist at SS: the fraction of thedefended prey clone is high but less than 100 per cent.This is followed by a substantial range of defence levelswhere the SS involves only the defended clone and thepredator. Eventually, for moderate to low levels ofdefence (0.3Rp1!p2 in figure 2), there are classicalconsumer–resource cycles (figure 2) with the predatorand only the defended clone. What happens whendefence is nearly ineffective ( p1zp2) is very sensitive tothe specifics of the trade-off curve. However, thequalitative pattern shown in figure 2 for high tomoderately effective defence is robust to changes inparameter values, and also to some qualitative changesin the model (i.e. the form of the predator functionalresponse, inclusion versus omission of predator agestructure ( Jones & Ellner 2004, 2007)).

The pattern predicted in figure 2 is somewhatcounter-intuitive, because it says that improveddefence by a defended clone is beneficial to theundefended clone. Specifically, an undefended clonecan persist in the system only when the defended cloneis nearly invulnerable to predation. What causes thispattern is indirect facilitation, a reversal of the familiar‘apparent competition’ scenario. Increased defence bythe defended prey (on its own) is bad for the predator,so it is good for undefended prey. As a prey lineageevolves better and better defence, eventually driving thepredator population to low numbers, it opens the wayfor a superior competitor that is not paying a cost(however, small) for anti-predator defence.

Some preliminary results (L. Becks, unpublisheddata; figure 3) from chemostat experiments withChlamydomonas reinhardii as the algal prey exhibit thepattern predicted by figure 2 with regard to EC versussteady-state coexistence. The Chlamydomonas systemhas the advantage that one prey defence trait, namelythe formation of clumps that are less easily consumedby the predator, is visible and easily quantified, so wecan track the dynamics of this trait along with thepopulation dynamics. We have shown that preyclumping reduces the ability of predators to consumethem, and that the tendency to clump is heritable(L. Becks, unpublished data). However, clumpingmay not be the only defence trait that these preycan evolve, so the degree of clumping may not bea complete indication of the level of prey defence.A direct indicator of the effectiveness of prey defence isthe relationship between prey abundance and predatorpopulation growth. The better defended the prey are,

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1

Phil. Trans. R. Soc. B

the lower the predator per capita rate of increase at anygiven prey abundance, and the higher the preyabundance required for the predator population toincrease rather than decrease.

In the experiment shown in figure 3a, the systemsettled to a SS with defended prey (as indicated by theoccurrence of clumping). The level of prey defence inthis experiment was such that the predator populationhad per capita rate of increase rz0 at algal density ofapproximately 105 mlK1 (days 45–58). The experimentshown in figure 3b was run under the same conditions,but was initialized with prey drawn from a populationthat had been exposed to predation for several months.The prey in this latter experiment developed far moreeffective defence than in the former experiment, as seenby the fact that (for example) at days 25 and 60 thepredator population was decreasing even though thealgal density was approximately 2!105 mlK1. Aspredicted in figure 2, stronger prey defence wasaccompanied by a shift from SS to EC. A completeanalysis of these experiments and additional replicateswill be presented elsewhere (Becks et al. in preparation).

(b) Systems with more than two prey types

The details of the trade-off between defence andcompetitive ability become important when morethan two prey types are initially present. Dependingon the shape of the trade-off curve, it may be possiblefor more than two prey types to coexist with thepredator. In that situation, the number of prey typespresent, the range of palatability between the least andmost defended types and the shape of the trade-offcurve together determine whether the system exhibits

–14

Page 6: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

0 0.2 0.4 0.6 0.8 1.0

1.0

0.4

0.8

1.4

1.8

(a) (b)

(c) (d )

(e) ( f )

palatability, p

k C(p

)/k C

(1)

b = 0.5

b = 1.0

b = 2.0

b = 4.0

time (days)

popu

latio

n de

nsiti

es

0.4

0.8

popu

latio

n de

nsiti

es

0.4

0.8

popu

latio

n de

nsiti

es

0.4

0.8

popu

latio

n de

nsiti

es

0.4

0.8

popu

latio

n de

nsiti

es

0

0

0

100 200 300 400time (days)

0

0

0

100 200 300 400

time (days)0 100 200 300 400

time (days)0 100 200 300 400

time (days)0

0

100 200 300 400

Figure 4. (a) Trade-off curves specified by equation (3.1). (b,c) Trajectories for a trade-off curve with bZ2, gZ1, for (b) NZ2,and (c) 32 initial prey types. Total prey are shown as dashed green curve and the predator is shown as solid red curve. A thin blackline denotes the mean predator density, which remains approximately 0.05 for all three scenarios in (a–c). The dilution rate isdZ0.7dK1, a setting for which we consistently observed cycles in our original experimental system (Jones & Ellner 2004). (d–f )Trajectories for a trade-off curve with bZ0.5, gZ1, for (d ) NZ2, (e) 4 and ( f ) 32 initial prey types. Total prey are shown ingreen and the predator is shown in red. A thin black line denotes the mean predator density, which is approximately 0.05, 0.07and 0.09, respectively, for the 2, 4 and 32 clone scenarios in (d–f ).

6 L. E. Jones et al. Rapid evolution in food webs

ARTICLE IN PRESS

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

707

708

709

710

711

712

713

714

715

716

717

718

719

720

721

722

723

724

725

726

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

746

747

748

749

750

751

752

753

754

755

756

757

758

759

760

761

762

763

764

765

766

767

768

stable limit cycles, transient cycling followed by acoexistence equilibrium of one or more types orexclusion of all but one prey genotype. The trade-offbetween defence against predation and competitiveability is modelled as follows:

kiðpiÞZk�

Cð1CgÞ

1Cgpbi

; i Z 1;2;.;N ; ð3:1Þ

where k�C is the half-saturation constant for the

completely undefended type; g is the cost of defence;and b sets the shape of the trade-off curve. Increaseddefence results in an increase in the half-saturationconstant ki( p), and thus a reduction in competitiveability relative to the undefended type in a predator-free environment. This curve (figure 4a) is only anexample of many trade-off formulations that produceconsistently similar results. For our purposes here,we set the minimum p-value as p1Z0.01, andpalatabilities pi take values between 0.01 and 1.Again, the mathematical model for the system isequation (2.1), but in this case with iZ1, 2, ., N.

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB –

Phil. Trans. R. Soc. B

For shape parameter bO1, the trade-off curve isconvex (figure 4a) for most settings of the costparameter g and favours extreme types (figure 5a–c).In the limit (as time t/N), the system behaves as atwo-clone system: given p1 is small, and any initialnumber of types arrayed on the trade-off curve, thereare stable EC for the dilution rate dZ0.7 (figure 4a–c).As defence drops ( p1 increases in value), there is first anarrow range of stable coexistence between the twoextreme types, p1 and pN, together with the predator.As with the two-prey system, there is then a substantialrange of defence levels where the SS involves only themost defended type ( p1) and the predator. As thedefence of the most defended type becomes moderate,similar to the two-prey system, there are thenconsumer–resource cycles. The small variation incycle period observed in figure 4b,c for differentnumbers of initial prey types is a transient phenom-enon, the duration of which increases as more preytypes are added to the system and the trade-off curvegets ‘crowded’: each subsequent prey type becomes

pp. 1–14

Page 7: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

1 2

–0.7

–0.4

–0.1(a) (b)

(c) (d )

(e) ( f )

log

(mea

n pr

ey f

requ

ency

)

1 2 3 4

–100

–40

0

–140

–20

–80

clone number

log

(mea

n pr

ey f

requ

ency

)

1 2

–0.7

–0.4

–0.1

1 2 3 4

–60

–30

0

log

(mea

n pr

ey f

requ

ency

)

log

(mea

n pr

ey f

requ

ency

)lo

g(m

ean

prey

fre

quen

cy)

log

(mea

n pr

ey f

requ

ency

)

–80

–40

0

1 7 14 22 30

1 7 14 22 30

clone number

Figure 5. Mean clonal frequencies (log scale, note differences in vertical scaling) for surviving prey types after 600 days at adilution rate dZ0.7dK1. (a–c) Assume that there were initially NZ(a) 2, (b) 4 and (c) 32 prey types arrayed on a trade-off curvewith shape parameter bZ2 and cost parameter gZ1. In this case, the surviving clones are the extremes (filled circles), and afterthe transient phase—which increases in length as prey types are added to the trade-off curve—the system behaves approximatelyas a two-clone system: it exhibits stable limit cycles at this dilution rate. (d–f ) Assume that there were initially NZ(d ) 2, (e) 4 and( f ) 32 prey types arrayed on a trade-off curve with shape parameter bZ0.5 and cost parameter gZ1. In this case, there is aninternal equilibrium with the most defended type and one or more very well-defended types (filled circles). The number ofcoexisting types increases with the number of clones and the concavity of the trade-off curve. Clones are indexed of the order ofincreasing palatability.

Rapid evolution in food webs L. E. Jones et al. 7

ARTICLE IN PRESS

769

770

771

772

773

774

775

776

777

778

779

780

781

782

783

784

785

786

787

788

789

790

791

792

793

794

795

796

797

798

799

800

801

802

803

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

822

823

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

849

850

851

852

853

854

855

856

857

858

859

860

861

862

863

864

865

866

867

868

869

870

871

872

873

874

875

876

877

878

879

880

881

882

883

884

885

886

887

888

889

890

891

892

893

894

895

896

more similar to those immediately adjacent to it on the

curve. Figure 5a–c shows surviving prey types for each

of two initial scenarios: NZ2, 4 or 32 prey types, when

bZ2. In each case, it is the extreme prey types (filled

circles) that survive in the presence of the predators,

with the most vulnerable prey type at very low densities

and the best defended at high densities.

For shape parameter b!1, the trade-off curve is

concave (figure 4a), and favours intermediate types.

Again assuming that the most defended type is nearly

invulnerable to predation, p1z0, as the number of prey

clones increases above two types, there is an initial phase

of transient EC followed by equilibrium during which

there is a stable coexistence of two or more of the best-

defended types. With an increase in prey types,

the transient phase becomes shorter and shorter,

and the dynamics are overall more stable with

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1

Phil. Trans. R. Soc. B

well-defended types dominating. In figure 4d–f, we

illustrate what happens as prey clonal diversity is

increased at the experimentally determined ‘cycling’

range of dZ0.7, and for p1z0 and trade-off shape

parameter bZ0.5: given NZ2 prey clones, the system

exhibits stable EC with a period of approximately

30 days and stable cycle amplitudes. As the clones are

increased to NZ4, the cycles rapidly shorten in period

and decline in amplitude with time, as the system goes to

equilibrium. By NZ32, the transient phase comprises

only one complete (evolutionary) cycle, and the system

is otherwise in equilibrium. Predator densities very

slowly increase with the number of initial clones, as the

dominant clone shifts to an intermediate type (e.g. see

figure 5, bZ0.5, NZ32). For further details and analysis

of coexistence equilibria in our experimentally para-

metrized model, please see Jones & Ellner (2004).

–14

Page 8: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

8 L. E. Jones et al. Rapid evolution in food webs

ARTICLE IN PRESS

897

898

899

900

901

902

903

904

905

906

907

908

909

910

911

912

913

914

915

916

917

918

919

920

921

922

923

924

925

926

927

928

929

930

931

932

933

934

935

936

937

938

939

940

941

942

943

944

945

946

947

948

949

950

951

952

953

954

955

956

957

958

959

960

961

962

963

964

965

966

967

968

969

970

971

972

973

974

975

976

977

978

979

980

981

982

983

984

985

986

987

988

989

990

991

992

993

994

995

996

997

998

999

1000

1001

1002

1003

1004

1005

1006

1007

1008

1009

1010

1011

1012

1013

1014

1015

1016

1017

1018

1019

1020

1021

1022

1023

1024

4. RESULTS II: DYNAMIC EFFECTS OF PREY ANDPREDATOR GENETIC VARIABILITYIn §3, we were able to give a rather detailed account ofthe effects of prey genetic diversity on communitydynamics. Of course, there is no a priori reason why thepredator in these systems should not also evolve.Indeed, we have shown that predator evolution doesoccur in our experimental rotifer–algal system and—just as prey evolution—can have strong effects on thestability of the observed predator–prey dynamics(Fussmann et al. 2003). We are only beginning toexplore how contemporary coevolution of predator andprey affects the dynamics. Because both predator andprey in our rotifer–algal system are clonal populations,genetic diversity can easily be introduced at bothtrophic levels simultaneously and the effects oncommunity dynamics can be studied. Given ourexperience with prey-evolution-only systems, weexpect, however, that understanding the resultingeco-evolutionary dynamics will require a detailedmathematical analysis that accounts for the multitudeof possible type-by-type interactions that occur withinand across trophic levels.

We here present a first step in this direction byformulating a predator–prey model that contains up tosix clones/genotypes within predator and prey levelsand allows for the interaction of each predator witheach prey type. So far, we are treating this model as anexploratory tool to understand the dynamical possibi-lities of this complex system, and intend to modelconcrete co-evolutionary predator–prey scenarios (i.e.rotifer and algal clones in the chemostat) in the future.The current model assumes logistic prey growth, amultispecies type-II functional response and is a systemof up to 12 differential equations (MZ6 prey and NZ6predator clones):

_xi Z xi ri 1KXMiZ1

xi =K

!KXNjZ1

aji yj

1CHj

" #; i Z1;.;M

and

_yj Z yj

Qj

1CHj

K dj

� �; j Z 1;.;N ;

ð4:1Þ

where xi and yi are the abundances of thesix prey and predator clones, respectively, and

HjZPM

iZ1bji xi ; QjZPM

iZ1 ajixi are measures of effec-

tive total prey abundance as perceived by predatorclone j. Parameters ri and dj are maximum prey clonegrowth rates and predator mortalities; parameters aji

and bji determine the saturation value and steepness ofthe functional responses for pairwise predator–preyclone interactions. The total abundance of all preyclones is regulated by the carrying capacity K. If noclonal diversity is present (MZNZ1), the systemsimplifies to the Rosenzweig–MacArthur model(Rosenzweig & MacArthur 1963). We parametrizedin agreement with values encountered in planktonpredator–prey systems (Fussmann & Heber 2002)that led to stable limit cycles for a system with a singleprey and a single predator type (figure 6a). Weintroduced clonal diversity by generating aji and bji

matrices with random uniform distribution (G5%)around the base values of aZ7.5 and bZ5.0. To see theeffects of heritable trait variation, we contrasted the

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB –

Phil. Trans. R. Soc. B

dynamics of a non-evolving (one-prey–one-predator)system using the base parameters, with the dynamicsof evolving multiple-clone systems with randomlygenerated parameters.

What are the effects on predator–prey dynamicswhen diversity exists at both the predator and preylevels? Numerical simulations of equation (4.1) resultin a wealth of dynamic scenarios but some interesting,repeatable patterns emerge.

(a) Selection processes among types are

important and affect the dynamics at the

species level

In all our simulations, only a subset of the six prey orpredator types persisted in the long run. Selectivesurvival of genotypes—due to natural selection—represents the evolutionary component of the eco-evolutionary dynamics we describe here. A frequentlyobserved outcome in a six-prey–one-predator system isthe selection of a single prey type that coexists with thepredator at stable equilibrium (figure 6b), whereas aone-prey–one-predator system with the prey having thebase parameter values displays limit cycle dynamicswith the predator (figure 6a). Stabilization occursbecause a prey type is selected that provides the highestgrowth rate in the presence of the predator, and at thesame time shifts the predator–prey pair to the stableside of the Hopf bifurcation. Selection of genotypes atboth predator and prey levels is the norm for the fullsix-prey–six-predator system, but very frequentlysystems of multiple prey and predator typesare selected that coexist on compensatory cycles(i.e. predator and prey types alternate regularly inrelative frequency and display oscillatory dynamicscharacterized by pairwise interactions of predator typeswith ‘their’ prey type; figure 6c,e). Figure 6e shows aninteresting example of coexistence of three predatorand three prey types. Each of the three distinguishablepredator–prey pair shows characteristic predator–preydynamics with the typical shift between predator andprey peaks. One predator–prey pair cycles much moreslowly than the other two, whose frequencies areapproximately five times higher. These examplesshow that when prey and predator are evolving intandem, the ‘cryptic cycles’ phenomenon can occur atboth trophic levels, with rapid dynamics at the clone-frequency level resulting in total species abundancesthat remain nearly constant.

(b) High type diversity is rare

Most frequently, systems with two predator and twoprey clones become selected. Complicated systemswith many types appear to be unable to coexist in thehomogeneous environment that this model represents.On the other hand, selection of just one clone of preyand predator each is not the most frequent outcome. Itseems that our six-by-six interaction model is able toreproduce the essential dynamical patterns expected inhomogeneous systems of much higher genetic diversitybecause they all reduce to systems of low to moderatediversity. Clonal genetic diversity in natural commu-nities is often much higher (e.g. De Meester et al.2006). This suggests that complex dynamics incomplex community networks are not a likely

pp. 1–14

Page 9: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

0

0

0.2

0.4

0.6

0.8

0.2

0

0.4

0.6

0.8

0.2

0

0.4

0.6

0.8

0.2

0.4

0.6

0.8(a) (b)

(c) (d )

(e) ( f )

(g) (h)

1000 2000time

3000 4000

0 1000 1500500 2000 2500 0 1000 1500500 2000 2500

0 1000 1500500 2000 25000 1000 1500500 2000 2500

0

0

40 6020 80 100 120 0 40 6020 80 100 120

0 1000 2000time

3000 4000

popu

latio

n de

nsity

popu

latio

n de

nsity

popu

latio

n de

nsity

popu

latio

n de

nsity

P P

P

P

P

P

P

P

Figure 6. Trajectories of predator–prey systems (equation (4.1)) with multiple prey (green) and predator (red; letter ‘P’)genotypes. (a) Benchmark system with one predator and one prey genotype. (b) Selection of a single prey type in a six-prey–one-predator system. (c) Selection of a two-prey–two-predator system in a six-prey–six-predator system. (d ) Species dynamicsof (c) (summation of predator and prey genotype abundances). (e) Selection of a three-prey–three-predator system in a six-prey–six-predator system. Note the two fast and one slowly cycling predator–prey pairs. ( f ) Species dynamics of (e).(g) Intermittent dynamics alternating between periods of dominance of single and multiple predator–prey pairs. (h) Speciesdynamics of (g). Parameters: riZ2.5, KZ1, djZ1, ajiZ7.5G5%, bjiZ5.0G5%. High-frequency oscillations appear ascontinuously filled areas in (d,f–h).

Rapid evolution in food webs L. E. Jones et al. 9

ARTICLE IN PRESS

1025

1026

1027

1028

1029

1030

1031

1032

1033

1034

1035

1036

1037

1038

1039

1040

1041

1042

1043

1044

1045

1046

1047

1048

1049

1050

1051

1052

1053

1054

1055

1056

1057

1058

1059

1060

1061

1062

1063

1064

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

1076

1077

1078

1079

1080

1081

1082

1083

1084

1085

1086

1087

1088

1089

1090

1091

1092

1093

1094

1095

1096

1097

1098

1099

1100

1101

1102

1103

1104

1105

1106

1107

1108

1109

1110

1111

1112

1113

1114

1115

1116

1117

1118

1119

1120

1121

1122

1123

1124

1125

1126

1127

1128

1129

1130

1131

1132

1133

1134

1135

1136

1137

1138

1139

1140

1141

1142

1143

1144

1145

1146

1147

1148

1149

1150

1151

1152

mechanism that maintains high levels of geneticdiversity in homogeneous natural systems.

(c) Genotype dynamics are wild, species and

community dynamics are calm

The increased complexity of the multi-predator–preysystem tends to lead to more complex and erratic

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1

Phil. Trans. R. Soc. B

dynamics at the genotype level (figure 6c,e,g),compared with the benchmark one-predator–one-preysystem (figure 6a). However, species dynamics(the sums of the abundances of predator and preytypes, respectively) become stabilized relative to thebenchmark system, as also observed by Vos et al. (2001)for a similar model system. This is partly due to

–14

Page 10: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

Q5

10 L. E. Jones et al. Rapid evolution in food webs

ARTICLE IN PRESS

1153

1154

1155

1156

1157

1158

1159

1160

1161

1162

1163

1164

1165

1166

1167

1168

1169

1170

1171

1172

1173

1174

1175

1176

1177

1178

1179

1180

1181

1182

1183

1184

1185

1186

1187

1188

1189

1190

1191

1192

1193

1194

1195

1196

1197

1198

1199

1200

1201

1202

1203

1204

1205

1206

1207

1208

1209

1210

1211

1212

1213

1214

1215

1216

1217

1218

1219

1220

1221

1222

1223

1224

1225

1226

1227

1228

1229

1230

1231

1232

1233

1234

1235

1236

1237

1238

1239

1240

1241

1242

1243

1244

1245

1246

1247

1248

1249

1250

1251

1252

1253

1254

1255

1256

1257

1258

1259

1260

1261

1262

1263

1264

1265

1266

1267

1268

1269

statistical averaging of time series (Cottingham et al.2001), but is primarily caused by the compensatorynature of the dynamics, i.e. the regular succession ofdifferent clones as opposed to their synchronousappearance and disappearance, and may lead toapparently steady-state dynamics at the species level(figure 6d, f ). This is another instance of crypticdynamics in which highly dynamic patterns of cyclicalselection of genotypes are hidden under a calm surfaceof species dynamics. Because this type of crypticdynamics can occur at very moderate levels of traitdiversity (parameters a and b diverging only up to 5%from their mean value), it is likely to play an importantrole in real communities, even when within-speciesdiversity is relatively low.

(d) Species dynamics may appear noisy

or display intermittency

Owing to the complexity of the dynamics, genotypeabundances never add up to smooth steady-state orcyclical species dynamics. Instead, equilibria appear‘noisy’ (figure 6d, f ) or the dynamics show intermittentpatterns. Intermittency occurs because periods domi-nated by multiple genotype interactions alternate withperiods of classical predator–prey oscillations (whenonly one prey and one predator genotype dominate;figure 6g,h). These simulations suggest that the noisyand intermittent patterns so frequently observed innatural communities are potentially due to intrinsic,cryptic genotype dynamics and are not necessarily theresult of changing environmental factors.

The patterns described here emerged from arelatively simple mathematical model in which we—somewhat artificially—generated the potential forevolutionary dynamics by setting initial levels of geneticdiversity of up to 12 genotypes. As such, weinvestigated the interaction between ecologicaldynamics and the dynamical process of naturalselection, but neglected processes that are able torestore genetic diversity. In natural communities,mutation and immigration of genotypes are suchprocesses, and future studies should consider theirimpact on eco-evolutionary dynamics. We suspect thatthe effects of pulsed mutation or immigration eventswill be understandable within the framework that wehave presented here because they just reset geneticdiversity to higher levels at fixed time intervals (as wedid at the start of our simulations). However,continuous rates of mutation or immigration at thetime scale of the ecological dynamics may lead todynamical outcomes not covered by our currentanalysis, and we see the inclusion of these processesas an important continuation of our work.

Q6

1270

1271

1272

1273

1274

1275

1276

1277

1278

1279

1280

5. DISCUSSIONAlthough the fields of ecology and evolutionary biologyare both over a century old, we are only beginning tounderstand how tightly coupled the processes at theheart of these disciplines truly are. Becoming some-body else’s dinner represents very strong naturalselection, and it is hard to imagine the predator–preyinteraction or any other strong interspecific interactionoccurring without direct evolutionary consequences.

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB –

Phil. Trans. R. Soc. B

But exactly that thinking is embedded in much of theecological theory underpinning environmental andnatural resource management. By studying somesimple model food webs based directly on experiments,we hope to start developing ecological theory for arapidly evolving biosphere.

We now return to the questions we posed in §1: howrapid contemporary evolution can affect the dynamicsof communities (question (i)), and then comparing thedynamic possibilities when genetic diversity is presentversus absent (question (ii)). Even our simplestexperimental systems (one predator and two preygenotypes) and the models developed to explain themgenerate a range of qualitatively different dynamics(§3). Changing the details of a trade-off between thecost of a defence, and its effectiveness can dramaticallyalter the identity of the surviving prey types, as canenvironmental factors (such as microcosm flow-through rate). Adding realistic, if theoretical, levels ofevolutionary complexity in the form of heritablevariation in ecologically relevant traits of both predatorand prey opens the door to even more possibilities (§4):selection of one prey type under some circumstances,and multiple types under others and dynamics varyingfrom equilibrium to chaos. Conversely, evolutionarydynamics can sometimes make for greater simplicity atthe species or community level, as complex dynamics atthe genotype level may lead to constant or nearlyconstant population abundances, instead of cycles, incoexisting populations (§§3 and 4).

A challenge posed by these findings is to developgeneral theory for linked eco-evolutionary dynamics,so that we can understand which biological conditionsgive rise to different ecological outcomes, and why.One approach that we are currently exploring is thetheory of slow–fast systems. This was used a decadeago (Khibnik & Kondrashov 1997) under the conven-tional assumption of slow evolutionary dynamicsrelative to fast ecological dynamics. To gain insightsinto possible effects of coupled and equally rapidevolutionary and ecological dynamics, we believe thatit will be useful to consider the opposite limit whereevolutionary change is fast relative to ecologicaldynamics (Cortez & Ellner in preparation). This isnot biologically impossible—it could happen if manybirths, balanced by nearly as many deaths, result inslow changes in total population size but allow rapidchange in genotype frequencies. Most importantly, itreduces model dimension and creates real possibilitiesfor mapping out all dynamic possibilities and under-standing when each of them can occur.

Because temporal dynamics provide informationabout underlying processes (e.g. Kendall et al. 2005;Ives et al. 2008), large-scale changes, such as con-sumer–resource cycles, are especially revealing aboutpotential effects of rapid contemporary evolution.Several types of natural large-scale ecological dynamicshave offered opportunities for studying rapid evolutionin the wild, including invasions (e.g. Lavergne &Molofsky 2007; Kinnison et al. 2008), range expansion(Reznick et al. 2008), periodic outbreaks of infectiousdiseases (e.g. Elderd et al. 2008), annual populationcycles driven by seasonality (Duffy & Sivars-Becker2007) and trait cycles (Sinervo & Lively 1996;

pp. 1–14

Page 11: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

Q7

Rapid evolution in food webs L. E. Jones et al. 11

ARTICLE IN PRESS

1281

1282

1283

1284

1285

1286

1287

1288

1289

1290

1291

1292

1293

1294

1295

1296

1297

1298

1299

1300

1301

1302

1303

1304

1305

1306

1307

1308

1309

1310

1311

1312

1313

1314

1315

1316

1317

1318

1319

1320

1321

1322

1323

1324

1325

1326

1327

1328

1329

1330

1331

1332

1333

1334

1335

1336

1337

1338

1339

1340

1341

1342

1343

1344

1345

1346

1347

1348

1349

1350

1351

1352

1353

1354

1355

1356

1357

1358

1359

1360

1361

1362

1363

1364

1365

1366

1367

1368

1369

1370

1371

1372

1373

1374

Sinervo et al. 2000). Systems at SS may also be shapedby rapid evolution, e.g. population stability may be anoutcome of the interaction between potentially rapidecological and evolutionary processes (e.g. Doebeli &Koella 1995; Zeineddine & Jansen 2005). But demon-strating what drives this outcome in natural settingsmay require experimental manipulations that arenecessarily high cost and high effort (e.g. Fischeret al. 2001), and whose interpretation has often beenproblematic because of inadvertent confoundingfactors resulting from large-scale interventions(Turchin 2003). Natural dynamic processes may thusoffer the best prospects for confronting theory withdata on consequences of rapid evolution, which iscritically important as the theory begins to develop.

Our most fundamental theoretical prediction is thatthe level of heritable variation in a trait affectingecological interactions is a bifurcation parameter:small gradual quantitative changes in unseen evolution-ary processes can cause large abrupt qualitativechanges at the population and community levels. Forthe practical issues of ecological management andforecasting that we posed in §1, it is already widelyrecognized that we need to consider possible effects ofevolutionary responses that are evoked by humanimpacts and interventions. Our results raise theconverse issue: we may also need to consider the effectsof evolutionary responses that are slowed or con-strained when genetic variation is reduced by humanimpacts, such as habitat loss or fragmentation.

1375

1376

1377

1378

1379

1380

1381

1382

1383

1384

1385

6. UNCITED REFERENCEHendry et al. (2008).

Our research has been supported primarily by the Andrew W.Mellon Foundation, the James S. McDonnell Foundationand the Natural Sciences and Engineering Research Councilof Canada. We thank Andrew Hendry and two anonymousreferees for their comments on the manuscript, and also themany Cornell undergraduates who assisted with datacollection.

1386

1387

1388

1389

1390

1391

1392

1393

1394

1395

1396

1397

1398

1399

1400

1401

1402

1403

1404

1405

1406

1407

1408

REFERENCESAbrams, P. A. 2005 ‘Adaptive dynamics’ vs. ‘adaptive

dynamics’. J. Evol. Biol. 18, 1162–1165. (doi:10.1111/j.1420-9101.2004.00843.x)

Abrams, P. A. & Matsuda, H. 1997 Prey adaptation as a causeof predator–prey cycles. Evolution 51, 1742–1750. (doi:10.2307/2410997)

Andersson, D. L. & Levin, B. R. 1999 The biological cost ofantibiotic resistance. Curr. Opin. Microbiol. 2, 489–493.(doi:10.1016/S1369-5274(99)00005-3)

Boraas, M. E., Searle, D. B. & Boxhorn, J. E. 1998Phagotrophy by a flagellate selects for colonial prey: apossible origin of multicellularity. Evol. Ecol. 12, 153–164.(doi:10.1023/A:1006527528063)

Cain, M. L., Bowman, W. D. & Hacker, S. D. 2008 Ecology.Sunderland, MA: Sinauer Associates, p. 621.

Cottingham, K. L., Brown, B. L. & Lennon, J. T. 2001Biodiversity may regulate the temporal variability ofecological systems. Ecol. Lett. 4, 72–85. (doi:10.1046/j.1461-0248.2001.00189.x)

Cousyn, C., De Meester, L., Colbourne, J. K., Brendonck, L.,Verschuren, D. & Volckaert, F. 2001 Rapid, localadaptation of zooplankton behavior to changes in

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1

Phil. Trans. R. Soc. B

predation pressure in the absence of neutral genetic

changes. Proc. Natl Acad. Sci. USA 98, 6256–6260.

(doi:10.1073/pnas.111606798)

De Meester, L., Vanoverbeke, J., De Gelas, K., Ortells, R. &

Spaak, P. 2006 Genetic structure of cyclic parthenogenetic

zooplankton populations—a conceptual framework. Arch.Hydrobiol. 167, 217–244. (doi:10.1127/0003-9136/2006/

0167-0217)

Doebeli, M. & Koella, J. C. 1995 Evolution of simple

population-dynamics. Proc. R. Soc. B 260, 119–125.

(doi:10.1098/rspb.1995.0068)

Dube, D., Kim, K., Alker, A. P. & Harvell, C. D. 2002 Size

structure and geographic variation in chemical resistance

of sea fan corals Gorgonia ventalina to a fungal pathogen.

Mar. Ecol. Prog. Ser. 231, 139–150. (doi:10.3354/

meps231139)

Duffy, M. A. & Sivars-Becker, L. 2007 Rapid evolution and

ecological host–parasite dynamics. Ecol. Lett. 10, 44–53.

(doi:10.1111/j.1461-0248.2006.00995.x)

Decaestecker, E., Gaba, S., Raeymaekers, J. A. M., Stoks, R.,

Van Kerckhoven, L., Ebert, D. & De Meester, L. 2007

Host–parasite ‘Red Queen’ dynamics archived in pond

sediment. Nature 450, 870–873. (doi:10.1038/

nature06291)

Eldred, B. D., Dushoff, J. & Dwyer, G. 2008 Host–pathogen

interactions, insect outbreaks, and natural selection for

disease resistance. Am. Nat. 172, 829–842. (doi:10.1086/

592403)

Ellner, S., Hairston Jr, N. G., Kearns, C. M. & Babaı, D.

1999 The roles of fluctuating selection and long-term

diapause in microevolution of diapause timing in a

freshwater copepod. Evolution 53, 111–122. (doi:10.

2307/2640924)

Ezard, T. H. G., Cote, S. D. & Pelletier, F. In press. Eco-

evolutionary dynamics: disentangling phenotypic,

environmental and population fluctuations. Phil. Trans.R. Soc. B 364. (doi:10.1098/rstb.2009.0006)

Ferriere, R. & Couvert, D. 2004 Evolutionary conservation

biology. Cambridge Studies in Adaptive Dynamics, vol. 4.

Cambridge, UK: Cambridge University Press.

Fischer, J. M., Klug, J. L., Ives, A. R. & Frost, T. M. 2001

Ecological history affects zooplankton community

responses to acidification. Ecology 82, 2984–3000.

Fussmann, G. F. & Heber, G. 2002 Food web complexity and

chaotic population dynamics. Ecol. Lett. 5, 394–401.

(doi:10.1046/j.1461-0248.2002.00329.x)

Fussmann, G. F., Ellner, S. P., Shertzer, K. W. & Hairston Jr,

N. G. 2000 Crossing the Hopf bifurcation in a live

predator–prey system. Science 290, 1358–1360. (doi:10.

1126/science.290.5495.1358)

Fussmann, G. F., Ellner, S. P. & Hairston Jr, N. G. 2003

Evolution as a critical component of plankton dynamics.

Proc. R. Soc. B 270, 1015–1022. (doi:10.1098/rspb.2003.

2335)

Fussmann, G. F., Loreau, M. & Abrams, P. A. 2007 Eco-

evolutionary dynamics of communities and ecosystems.

Funct. Ecol. 21, 465–477. (doi:10.1111/j.1365-2435.

2007.01275.x)

Gagneux, S., Long, C. D., Small, P. M., Van, T., Schoolnik,

G. K. & Bohannan, B. J. M. 2006 The competitive cost of

antibiotic resistance in Mycobacterium tuberculosis. Science

312, 1944–1946. (doi:10.1126/science.1124410)

Grant, P. R. & Grant, B. R. 2002 Unpredictable evolution in a

30-year study of Darwin’s finches. Science 296, 707–711.

(doi:10.1126/science.1070315)

Hairston Jr, N. G. & De Meester, L. 2008 Daphnia

paleogenetics and environmental change: deconstructing

the evolution of plasticity. Int. Rev. Hydrobiol. 93,

578–592. (doi:10.1002/iroh.200811057)

–14

Page 12: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

12 L. E. Jones et al. Rapid evolution in food webs

ARTICLE IN PRESS

1409

1410

1411

1412

1413

1414

1415

1416

1417

1418

1419

1420

1421

1422

1423

1424

1425

1426

1427

1428

1429

1430

1431

1432

1433

1434

1435

1436

1437

1438

1439

1440

1441

1442

1443

1444

1445

1446

1447

1448

1449

1450

1451

1452

1453

1454

1455

1456

1457

1458

1459

1460

1461

1462

1463

1464

1465

1466

1467

1468

1469

1470

1471

1472

1473

1474

1475

1476

1477

1478

1479

1480

1481

1482

1483

1484

1485

1486

1487

1488

1489

1490

1491

1492

1493

1494

1495

1496

1497

1498

1499

1500

1501

1502

1503

1504

1505

1506

1507

1508

1509

1510

1511

1512

1513

1514

1515

1516

1517

1518

1519

1520

1521

1522

1523

1524

1525

1526

1527

1528

1529

1530

1531

1532

1533

1534

1535

1536

Hairston Jr, N. G. & Walton, W. E. 1986 Rapid evolution of a

life-history trait. Proc. Natl Acad. Sci. USA 83,

4831–4833. (doi:10.1073/pnas.83.13.4831)

HairstonJr,N.G.,Lampert,W.,Caceres,C.E.,Holtmeier,C.L.,

Weider, L. J., Gaedke, U., Fischer, J. M., Fox, J. A. & Post,

D. M. 1999 Lake ecosystems—rapid evolution revealed by

dormant eggs. Nature 401, 446. (doi:10.1038/46731)

Hairston Jr, N. G., Ellner, S. P., Geber, M. A., Yoshida, T. &

Fox, J. A. 2005 Rapid evolution and the convergence of

ecological and evolutionary time. Ecol. Lett. 8, 1114–1127.

(doi:10.1111/j.1461-0248.2005.00812.x)

Heath, D. D., Heath, J. W., Bryden, C. A., Johnson, R. M. &

Fox, C. W. 2003 Rapid evolution of egg size in captive

salmon. Science 299, 1738–1740. (doi:10.1126/science.

1079707)

Hendry, A. P. & Kinnison, M. T. 1999 The pace of modern

life: measuring rates of contemporary microevolution.

Evolution 53, 1637–1653. (doi:10.2307/2640428)

Hendry, A. P., Farrugia, T. & Kinnison, M. T. 2008 Human

influences on rates of phenotypic change in wild animal

populations. Mol. Ecol. 17, 20–29. (doi:10.1111/j.1365-

294X.2007.03428.x)

Ives, A. R., Einarsson, A., Jansen, V. A. A. & Gardarsson, A.

2008 High-amplitude fluctuations and alternative dyna-

mical states of midges in Lake Myvatn. Nature 452, 84–87.

(doi:10.1038/nature06610)

Johnson, M. T. J., Lajeunesse, M. J. & Agrawal, A. 2006

Additive and interactive effects of plant genotypic diversity

on arthropod communities and plant fitness. Ecol. Lett. 9,

24–34.

Johnson, M. T. J., Vellend, M. & Stinchcombe, J. R. In press.

Evolution in plant populations as a driver of ecological

changes in arthropod communities. Phil. Trans. R. Soc. B

364. (doi:10.1098/rstb.2008.0334)

Jones, L. E. & Ellner, S. P. 2004 Evolutionary tradeoff and

equilibrium in an aquatic predator–prey system. Bull.

Math. Biol. 66, 1547–1573. (doi:10.1016/j.bulm.2004.02.

006)

Jones, L. E. & Ellner, S. P. 2007 Effects of rapid prey

evolution on predator–prey cycles. J. Math. Biol. 55,

541–573. (doi:10.1007/s00285-007-0094-6)

Kendall, B. E., Ellner, S. P., McCauley, E., Wood, S. N.,

Briggs, C. J., Murdoch, W. W. & Turchin, P. 2005

Population cycles in the pine looper moth Bupalus

piniarius: dynamical tests of mechanistic hypotheses.

Ecol. Monogr. 75, 259–276. (doi:10.1890/03-4056)

Khibnik, A. I. & Kondrashov, A. S. 1997 Three mechanisms

of Red Queen dynamics. Proc. R. Soc. B 264, 1049–1056.

(doi:10.1098/rspb.1997.0145)

Kinnison, M. T. & Hairston Jr, N. G. 2007 Eco-evolutionary

conservation biology: contemporary evolution and the

dynamics of persistence. Funct. Ecol. 21, 444–454.

(doi:10.1111/j.1365-2435.2007.01278.x)

Kinnison, M. T. & Hendry, A. P. 2001 The pace of modern

life II: from rates of contemporary microevolution to

pattern and process. Genetica 112–113, 45–164.

Kinnison, M. T., Unwin, M. J. & Quinn, T. P. 2008 Eco-

evolutionary vs. habitat contributions to invasion in

salmon: experimental evaluation in the wild. Mol. Ecol.

17, 405–414. (doi:10.1111/j.1365-294X.2007.03495.x)

Lavergne, S. & Molofsky, J. 2007 Increased genetic variation

and evolutionary potential drive the success of an invasive

grass. Proc. Natl Acad. Sci. USA 104, 3883–3888. (doi:10.

1073/pnas.0607324104)

Lenski, R. E. & Travisano, M. 1994 Dynamics of adaptation

and diversification: a 10,000-generation experiment with

bacterial populations. Proc. Natl Acad. Sci. USA 91,

6808–6814. (doi:10.1073/pnas.91.15.6808)

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB –

Phil. Trans. R. Soc. B

Meyer, J. R. & Kassen, R. 2007 The effects of competition andpredation on diversification in the model adaptive radiation.Nature 446, 432–435. (doi:10.1038/nature05599)

Meyer, J. R., Ellner, S. P., Hairston Jr, N. G., Jones, L. E. &Yoshida, T. 2006 Prey evolution on the time scale ofpredator–prey dynamics revealed by allele-specific quan-titative PCR. Proc. Natl Acad. Sci. USA 103,10 690–10 695. (doi:10.1073/pnas.0600434103)

Molles Jr, M. C. 2008 Ecology: concepts, applications, 4th edn.New York, NY: McGraw Hill.

Olsen, E. M., Heino, M., Lilly, G. R., Morgan, M. J., Brattey, J.,Ernande, B. & Dieckmann, U. 2004 Maturation trendsindicative of rapid evolution preceded the collapse ofnorthern cod. Nature 428, 932–935. (doi:10.1038/nature02430)

Palkovacs, E. P., Marshall, M. C., Lamphere, B. A., Lynch,B. R., Weese, D. J., Fraser, D. F., Reznick, D. N., Pringle,C. M. & Kinnison, M. T. In press. Experimentalevaluation of evolution and coevolution as agents ofecosystem change in Trinidadian streams. Phil. Trans. R.Soc. B 364. (doi:10.1098/rstb.2009.0016)

Palumbi, S. R. 2001 The evolution explosion: how humanscause rapid evolutionary change. New York, NY: Nortonp. 277.

Pelletier, F., Clutton-Brock, T., Pemberton, J., Tuljapurkar,S. & Coulson, T. 2007 The evolutionary demography ofecological change: linking trait variation and populationgrowth. Science 315, 1571–1574. (doi:10.1126/science.1139024)

Post, D. M. & Palkovacs, E. P. In press. Eco-evolutionaryfeedbacks in community and ecosystem ecology:interactions between the ecological theatre and theevolutionary play. Phil. Trans. R. Soc. B 364. (doi:10.1098/rstb.2009.0012)

Reznick, D. N. & Ghalambor, C. K. 2001 The populationecology of contemporary adaptations: what empiricalstudies reveal about the conditions that promote adaptiveevolution. Genetica 112, 183–198. (doi:10.1023/A:1013352109042)

Reznick, D. N., Shaw, F. H., Rodd, F. H. & Shaw, R. G. 1997Evaluation of the rate of evolution in natural populationsof guppies (Poecilia reticulata). Science 275, 1934–1937.(doi:10.1126/science.275.5308.1934)

Reznick, D. N., Ghalambor, C. K. & Crooks, K. 2008Experimental studies of evolution in guppies: a model forunderstanding the evolutionary consequences of predatorremoval in natural communities. Mol. Ecol. 17, 97–107.(doi:10.1111/j.1365-294X.2007.03474.x)

Rosenzweig, M. L. & MacArthur, R. H. 1963 Graphicalpresentation and stability conditions of predator–preyinteractions. Am. Nat. 97, 209–223. (doi:10.1086/282272)

Shertzer, K. W., Ellner, S. P., Fussmann, G. F. & Hairston Jr,N. G. 2002 Predator–prey cycles in an aquatic microcosm:testing hypotheses of mechanism. J. Anim. Ecol. 71,802–815. (doi:10.1046/j.1365-2656.2002.00645.x)

Sinervo, B. & Lively, C. M. 1996 The rock-paper-scissorsgame and the evolution of alternative male strategies.Nature 380, 240–243. (doi:10.1038/380240a0)

Sinervo, B., Svensson, E. & Comendant, T. 2000 Densitycycles and an offspring quantity and quality game drivenby natural selection. Nature 406, 985–988. (doi:10.1038/35023149)

Strauss, S. Y., Lau, J. A., Schoener, T. W. & Tiffin, P. 2008Evolution in ecological field experiments: implications foreffect size. Ecol. Lett. 11, 199–207. (doi:10.1111/j.1461-0248.2007.01128.x)

Swain, D. P., Sinclair, A. F. & Mark Hanson, J. 2007Evolutionary response to size-selective mortality in anexploited fish population. Proc. R. Soc. B 274, 1015–1022.(doi:10.1098/rspb.2006.0275)

pp. 1–14

Page 13: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

Rapid evolution in food webs L. E. Jones et al. 13

ARTICLE IN PRESS

1537

1538

1539

1540

1541

1542

1543

1544

1545

1546

1547

1548

1549

1550

1551

1552

1553

1554

1555

1556

1557

1558

1559

1560

1561

1562

1563

1564

1565

1566

1567

1568

1569

1570

1571

1572

1573

1574

1575

1576

1577

1578

1579

1580

1581

1582

1583

1584

1585

1586

1587

1588

1589

1590

1591

1592

1593

1594

1595

1596

1597

1598

1599

1600

1601

1602

1603

1604

1605

1606

1607

1608

1609

1610

1611

1612

1613

1614

1615

Thompson, J. N. 1998 Rapid evolution as an ecologicalprocess. Trends Ecol. Evol. 13, 329–332. (doi:10.1016/S0169-5347(98)01378-0)

Turchin, P. 2003 Complex population dynamics. Princeton, NJ:Princeton University Press.

Urban, M. C. et al. 2008 The evolutionary ecology ofmetacommunities. Trends Ecol. Evol. 23, 311–317.(doi:10.1016/j.tree.2008.02.007)

Vos,M., MorenoBerrocal,S., Karamaouna,F.,Hemerick, L. &Vet, L. E. M. 2001 Plant-mediated indirect effectsand the persistence of parasitoid–herbivore communities.Ecol. Lett. 4, 38–45. (doi:10.1046/j.1461-0248.2001.00191.x)

Whitham, T. G. et al. 2006 A framework for community andecosystem genetics: from genes to ecosystems. Nat. Rev.Genet. 7, 510–523. (doi:10.1038/nrg1877)

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1

Phil. Trans. R. Soc. B

Yoshida, T., Jones, L. E., Ellner, S. P., Fussmann, G. F. &Hairston Jr, N. G. 2003 Rapid evolution drives ecologicaldynamics in a predator–prey system. Nature 424, 303–306.(doi:10.1038/nature01767)

Yoshida, T., Hairston Jr, N. G. & Ellner, S. P. 2004Evolutionary trade-off between defence against grazingand competitive ability in a simple unicellular alga,Chlorella vulgaris. Proc. R. Soc. B 271, 1947–1953.(doi:10.1098/rspb.2004.2818)

Yoshida, T., Ellner, S. P., Jones, L. E., Bohannan, B. J. M.,Lenski, R. E. & Hairston Jr, N. G. 2007 Cryptic populationdynamics: rapid evolution masks trophic interactions. PLoSBiol. 5, 1868–1879. (doi:10.1371/journal.pbio.0050235)

Zeineddine, M. & Jansen, V. A. A. 2005 The evolution ofstability in a competitive system. J. Theor. Biol. 236,208–215. (doi:10.1016/j.jtbi.2005.03.004)

–14

1616

1617

1618

1619

1620

1621

1622

1623

1624

1625

1626

1627

1628

1629

1630

1631

1632

1633

1634

1635

1636

1637

1638

1639

1640

1641

1642

1643

1644

1645

1646

1647

1648

1649

1650

1651

1652

1653

1654

1655

1656

1657

1658

1659

1660

1661

1662

1663

1664

Page 14: Rapid contemporary evolution and clonal food web dynamics · Rapid contemporary evolution and clonal food web dynamics Laura E. Jones1, Lutz Becks1,†, Stephen P. Ellner1, Nelson

14 L. E. Jones et al. Rapid evolution in food webs

ARTICLE IN PRESS

1665

1666

1667

1668

1669

1670

1671

1672

1673

1674

1675

1676

1677

1678

1679

1680

1681

1682

1683

1684

1685

1686

1687

1688

1689

1690

1691

1692

1693

1694

1695

1696

1697

1698

1699

1700

1701

1702

1703

1704

1705

1706

1707

1708

1709

1710

1711

1712

1713

1714

1715

1716

1717

1718

1719

1720

1721

1722

1723

1724

1725

1726

1727

1728

1729

1730

1731

1732

1733

1734

1735

1736

1737

1738

1739

1740

1741

1742

1743

1744

1745

1746

1747

1748

1749

1750

1751

1752

1753

1754

1755

1756

1757

Author Queries

JOB NUMBER: 20090004

JOURNAL: RSTB

Q1 Reference Duffy et al. (2007) has been cited in text but

not provided in the list. Please supply reference details

or delete the reference citation from the text.

Q2 We have changed ‘figure2e versus figure 2c’ to

‘figure1e versus figure 1c’ in the sentence ‘In addition,

the coefficient. .’ Please check and approve.

Q3 Please give the year during which Becks (unpublished

data) were generated.

Q4 Reference Becks et al. (in preparation) has been cited

in text but not provided in the list. Please supply

reference details or delete the reference citation from

the text.

Q5 Reference Cortez & Ellner (in preparation) has been

cited in text but not provided in the list. Please supply

reference details or delete the reference citation from

the text.

Q6 Reference Elderd et al. (2008) has been cited in text

but not provided in the list. Please supply reference

details or delete the reference citation from the text.

Q7 Reference Hendry et al. (2008) is provided in the list

but not cited in the text. Please supply citation details

or delete the reference from the reference list.

RSTB 20090004—1/3/2009—15:50—JMANOHAR—329484—XML RSB – pp. 1–14

1758

1759

1760

1761

1762

1763

1764

1765

1766

1767

1768

1769

1770

1771

1772

1773

1774

1775

1776

1777

1778

1779

1780

1781

1782

1783

1784

1785

1786

1787

1788

1789

1790

1791

1792

Phil. Trans. R. Soc. B