Biophysical models of small pelagic fish

7 downloads 287 Views 702KB Size Report
of pelagic fish (Bakun, 1996). The results of Lett et al. (2006) ..... how water releases from dams would affect fish downstream (Acreman and Dunbar, 2004). Use.
1

Chapter 6

2 3

Biophysical models of small pelagic fish

4 5

Christophe Letta, Kenneth A. Roseb, Bernard A. Megreyc

6 7

a

8

Oceanography Department, Rondebosch, 7701, South Africa, [email protected]

Institut de Recherche pour le Développement, UR ECO-UP, University of Cape Town,

9 10

b

11

LA 70803, USA, [email protected]

Department of Oceanography and Coastal Sciences, Louisiana State University, Baton Rouge,

12 13

c

14

Fisheries Science Center, Seattle, WA 98115, USA, [email protected]

National Oceanic and Atmospheric Administration, National Marine Fisheries Service, Alaska

15

1

15

Biophysical models of fish early life stages are used to simulate the dynamics of eggs and larvae

16

in virtual marine environments. Environments are characterised by temporally dynamic three-

17

dimensional fields of physical (e.g., current velocities, temperature) and biogeochemical (e.g.,

18

phytoplankton and zooplankton concentrations) variables provided by hydrodynamic or

19

hydrodynamic-biogeochemical coupled models. These fields are used as inputs to fish models

20

that simulate the dynamics of eggs and larvae including their transport, growth, mortality and

21

behavioural processes (Figure 1). Biophysical models of early life stages are ready for

22

investigating the effects of climate change on egg and larval growth, mortality, and spatial

23

distributions. The quality of the predictions of the biophysical models will depend on sufficient

24

accuracy in the physical and biogeochemical models under climate change scenarios, whether the

25

spatial and temporal scales of the physical and biogeochemical model predictions are appropriate

26

to be used as inputs to biophysical fish models, and the degree to which the growth (including

27

feeding), mortality (including predation), and behaviour-based movement processes represented

28

in the biophysical model have adequate species-specific and site-specific details.

29 30

Many questions related to the effects of climate change on fish population dynamics (e.g.,

31

sustainable harvest) require the inclusion of juvenile and adult life stages in the biophysical

32

models. Unlike for eggs and larvae for which physics plays a dominant role in their movement,

33

behaviour plays an important role in the movement of juvenile and adult fish. Biophysical models

34

that include the adult life stages of fish is an area of high research interest but will likely be used

35

for general analysis of climate effects or application to a few, extremely well-studied species and

36

locations. The uncertainty in how to model movement will limit the development of a generally

37

agreed upon modelling approach that has helped advance the early life stage models. Continued

38

efforts on adult models is necessary in order to get to “end-to-end” (physics to fish), full life cycle

39

models capable of addressing the long-term consequences of climate change on fish and fisheries.

40

2

40

Abstract

41 42

The objective of this chapter is to provide an overview of biophysical models of marine fish

43

populations, with a particular focus on those applied, or potentially applicable, for examining the

44

consequences of climate change on small pelagic fish species. We focus on models that include

45

physics and are therefore spatially-explicit, and review models under the categories of those that

46

include lower trophic level dynamics (NPZ), early life stages of fish (eggs and larvae), and

47

juvenile and adult stages. We first give an overview of the methods that are used to represent

48

transport, growth, mortality, and behaviour in biophysical models of early life stages. Second, we

49

detail several case studies of such models, focusing on those applied to anchovy and sardine in

50

SPACC regions and those involving small pelagic fish in “non-SPACC” regions. Some questions

51

related to climate change require models that include juveniles and adults. Models that include

52

juveniles and adults differ from the early life stage models in the important role played by

53

behaviour in fish movement. We briefly discuss several approaches used for modelling

54

behavioural movement of fish, and then summarize several case studies of biophysical models

55

that include adults that are relevant, or potentially relevant, to small pelagic species. Finally, we

56

conclude with a discussion of the potential use of biophysical models of early life stages and

57

adults for investigating some of the issues associated with forecasting the effects of climate

58

change on small pelagic fish species.

59 60

1 Introduction

61 62

Understanding and forecasting the effects of climate change on small pelagic fish involves

63

coupling the physics and lower trophic level dynamics to the growth, mortality, reproduction, and

64

movement processes of key life stages that govern fish recruitment and population dynamics.

65

Fish exhibit large changes in body weight, and often dramatic changes in body form, habitat

66

usage, and diet, with ontogenetic development through the egg, larval, juvenile, and adult life

67

stages. Early life stages (eggs and larvae) are heavily influenced by advective processes, which

68

determine where they go in the system and to what environmental conditions they are exposed.

69

There are many models that have coupled physics with eggs and larvae dynamics, and examined

3

70

how physical transport under different conditions can affect the growth and mortality rate

71

experienced by individuals during these early life stages (Werner et al., 2001; Runge et al., 2005;

72

Miller, in press). The “Workshop on advancements in modeling physical-biological interactions

73

in fish early-life history: recommended practices and future directions” held on 3-5 April 2006 in

74

Nantes, France (co-chairs: A. Gallego, E. North and P. Petitgas), attracted more than 50

75

participants from 14 countries, an indication of the international interest in the topic. There are

76

now modelling tools that allow designing simulations coupling physics with ichthyoplankton

77

dynamics (e.g., Ichthyop, Lett et al., submitted, http://www.eco-up.ird.fr/projects/IBM.htm).

78 79

Many questions posed about climate effects on fish and fisheries can be only partially answered

80

by models that predict growth and survivorship to the larval or early juvenile life stage.

81

Recruitment for some species may not be determined until later in the life cycle than is simulated

82

with these early life stage models. Furthermore, some questions require simulating the effects of

83

climate on adults (e.g., how will climate affect migration, spatial distributions and spawning

84

success), and other questions require multi-generational simulations that include all life stages in

85

order to forecast the long-term consequences of climate effects (e.g., climate effects on

86

sustainable harvest levels). The coupling of physics and the lower trophic levels to models that

87

include the adult life stages is much less common than models restricted to early life stages.

88

Biophysical models that include adult life stages are being increasingly considered and their use,

89

especially with the need to investigate the effects of climate and fishing in marine ecosystems,

90

will accelerate over the next decade (Travers et al., submitted).

91 92

The objective of this chapter is to provide an overview of biophysical models of fish, with a

93

particular focus on those applied, or potentially applicable, for examining the consequences of

94

climate change on small pelagic species. We consider models of early life stages and models that

95

include adults. Our review is not intended to be comprehensive, and others have previously

96

reviewed biophysical models of marine populations (Werner et al., 2001; Runge et al., 2005;

97

Miller, in press). We focus our review on models of fish that include physics and possibly lower

98

trophic level dynamics; inclusion of physics implies the models must be spatially-explicit. We

99

divided these models into two general categories: single-species individual-based models (IBMs)

100

of fish early life history and models of adult stages (either adults only, or full life cycle that

4

101

include all life stages). In the early life stage IBMs, the abiotic environment is described with

102

outputs from a hydrodynamic model, or by a hydrodynamic model coupled to a biogeochemical

103

(e.g., NPZ) model. Werner et al. (2001) classified IBMs in which the physics is used for transport

104

but the biotic (NPZ) environment is absent or represented by static prey fields derived from field

105

data as “hydrodynamics and simple behaviors” or “hydrodynamics and static prey”. Building on

106

the classification proposed by Werner et al., we also include in our review IBMs that use outputs

107

of biogeochemical models as a dynamic representation of the prey for the fish, and these we term

108

“hydrodynamics and dynamic prey”. The adult models are much less common and, while we

109

focus on models that are spatially explicit and coupled to hydrodynamic and biogeochemical

110

models, we also include models that do not fulfil all of the criteria. We included some adult

111

models because they could be adapted to examine questions about climate change effects that

112

require the inclusion of adult life stages.

113 114

This chapter is organized as follows. First, we give an overview of the methods used to represent

115

transport, growth, mortality, and behaviour of individuals in biophysical models of early life

116

stages. Second, we detail several case studies of such models, organized by models of anchovy

117

and sardine in SPACC regions (first in the Benguela upwelling system, then in other major

118

upwelling systems), and for small pelagic fish in general in “non-SPACC” regions. We then turn

119

to models that include adults. Because fish behaviour comes much more into play when adult

120

stages are considered, advective transport is no longer sufficient and we discuss methods used to

121

represent movement. We then detail several case studies of biophysical models that include adults

122

that are relevant, or potentially relevant, to small pelagic species and climate change. Finally, we

123

conclude with a discussion of the potential use of these biophysical models to investigate some of

124

the issues raised in the context of climate change.

125 126

2 Biophysical models of early life stages

127 128

2.1 Modelling the fish egg and larval environment

129

5

130

2.1.1 The abiotic environment

131 132

The abiotic environment for the fish eggs and larvae is usually provided by simulations of

133

hydrodynamic models. In some studies, field data were directly used to generate the needed

134

physics for model simulations (e.g., Heath et al., 1998; Rodríguez, 2001; Santos et al., 2004). The

135

hydrodynamic models were generally used to provide temporally dynamic, 3-dimensional fields

136

of physical variables such as current velocities, temperature, and salinity.

137 138

2.1.2 The biotic environment

139 140

The use of biogeochemical models or field data to generate the biotic environment for the fish

141

eggs and larvae has received less consideration. Most applications focus on the prey field for

142

influencing the feeding and growth rates of the fish larvae. Prey fields have been generated as

143

input to the early life stage models using biogeochemical models (e.g., Hinckley, 1999; Koné,

144

2006), interpolated from field data (e.g., Hinrichsen et al., 2002; Lough et al., 2006), or assessed

145

using remotely-sensed data (e.g., Bartsch and Coombs, 2004). On the other hand, use of models

146

or data to specify spatially and temporally varying predation on fish eggs and larvae is rare (but

147

see Suda and Kishida, 2003; Suda et al., 2005; Vikebø et al., in press). Predation is usually

148

considered part of the total mortality rate, with the changing vulnerability of early life stages

149

represented by mortality rates being stage-specific or size-dependent.

150 151

2.2 Modelling fish egg and larval dynamics

152 153

Most biophysical models of the early life stages are individual-based models (Lagrangian),

154

although a few models used an Eulerian approach (e.g., Zakardjian et al., 2003, who applied their

155

model to zooplankton dynamics). There are a variety of different methods for coupling a

156

hydrodynamic model to an individual-based model of eggs and larvae (Hermann et al., 2001).

157

Most common is to run the hydrodynamic model, store the outputs, and then use the outputs as

158

inputs to the fish model (Figure 1).

6

159 160

Figure 1

161 162

2.2.1 Transport

163 164

v Transport of individuals in biophysical models consists in updating the individuals’ position x

165

using the following equation:

166 r r r dx / dt = u + u ′

(1)

167

169

r The deterministic advection term u is typically derived from spatial and temporal interpolations r of the flow fields provided by the hydrodynamic model. A stochastic term u ′ is often added to

170

take into account the small-scale fluctuations in the currents that are lost due to the grid resolution

171

of the hydrodynamic model and temporal averaging of the hydrodynamic model outputs. The

172

stochastic term is often represented using an ad-hoc diffusion term (i.e., a random walk process)

173

or using a random displacement process if a spatially non-uniform diffusivity is used (North et

174

al., 2006; Huret et al., in press). Recent theoretical results using the dispersal characteristics of

175

drifters show promise for objectively specifying the stochastic term (Haza et al., in press).

176

Equation (1) is numerically integrated using schemes such as a forward Euler or a Runge-Kutta.

177

Ådlandsvik (abstract 1 ) has recently proposed several standard test cases to evaluate the advection

178

scheme used in biophysical models. Some studies have specifically attempted to “validate” the

179

flow fields and the transport scheme used by comparing trajectories of virtual and real drifters

180

(Gutiérrez et al., 2004; Thorpe et al., 2004; Edwards et al., 2006; Fach and Klinck, 2006).

168

181 182

2.2.2 Growth

183

1

Ådlandsvik, B. The particle-tracking method for transport modelling. Workshop on advancements in modeling physical-biological interactions in fish early-life history: recommended practices and future directions. Nantes, France, April 3-5, 2006.

7

184

A general relationship between the attributes of individuals (e.g., length L and age) and

185

environmental variables (e.g., temperature T and food biomass F ) was proposed by Heath and

186

Gallego (1997) to simulate growth:

187 t = age

Lage =

∫ f (age) f 1

2

(T ) f 3 ( F )dt

(2)

t =0

188 189

In most biophysical models, the growth algorithm used various simplified versions of equation

190

(2) in which stage durations (e.g., Miller et al., 1998), length (e.g., Fox et al., 2006), or weight

191

(e.g., Vikebø et al., 2005) were functions of temperature only. When the effect of prey fields on

192

growth was considered, relatively more complex bioenergetics sub-models were developed (e.g.,

193

Hinckley, 1999; Megrey and Hinckley, 2001; Hinrichsen et al., 2002; Lough et al., 2005). These

194

bioenergetics models allow for weight and temperature effects on consumption and the loss terms

195

(respiration, egestion, excretion), and use the prey field information to determine the actual

196

realized consumption rate. There is an extensive set of bioenergetics models for fish, many of

197

which use the same notation and formulations in what is termed the Wisconsin model that allows

198

for easy inter-specific and inter-life stage comparisons (Ney, 1993; Hanson et al., 1997).

199 200

2.2.3 Mortality

201 202

Mortality has been represented in early life stage models in a variety of ways, including constant

203

rates (e.g., Brown et al., 2005; Tilburg et al., 2005), and depending on life stage (e.g., Miller et

204

al., 1998), length (e.g., Hinckley, 1999), weight (e.g., Brickman et al., 2001), temperature (e.g.,

205

Mullon et al., 2003), and growth rate (e.g., Hinrichsen et al., 2002; Bartsch and Coombs, 2004).

206

Suda and colleagues (Suda and Kishida 2003; Suda et al., 2005) is one of the few examples that

207

included inter-specific effects and density dependent effects on mortality. Vikebø et al. (in press)

208

considered predation rates on larvae as functions of their size and light intensity. The high

209

mortality rates of early life stage IBMs can result in numerical problems because many

210

individuals need to be followed in order to obtain enough survivors to allow analysis of the

211

results. One method for dealing with the high mortality problem is to use the concept of super-

8

212

individuals (Scheffer et al., 1995), in which each simulated individual is assumed to represent

213

some number of identical population individuals (e.g., Hinckley, 1999; Megrey and Hinckley,

214

2001; Allain, 2004; Bartsch and Coombs, 2004). Rather than mortality resulting in the removal of

215

model individuals, the worth of each super-individual is decreased based on the mortality rate.

216

Models that use super-individuals can therefore use a constant, and a priori determined, number

217

of model individuals throughout their simulations. Caution is needed in using super-individuals to

218

ensure numerical accuracy of model simulations. When processes are density-dependent or the

219

introduction of new individuals occurs over an extended period of time or space, proper balancing

220

of how new super-individuals are introduced (number, initial population number they represent,

221

timing and spatial location of their introduction) is needed to ensure accurate model results.

222 223

2.2.4 Behaviour

224 225

Egg buoyancy schemes have been included in determining the vertical position of individuals.

226

Typically, a vertical velocity term is used that is assumed to be proportional to the difference

227

between egg density and water density. Egg densities have been assumed to be constant (e.g.,

228

Parada et al., 2003; North et al., 2005), stage-dependent (e.g., Hinckley, 1999), and time-

229

dependent (e.g., Brickman et al., 2001). Because growth, mortality, and horizontal advection are

230

all functions of depth, a key behavioural trait for drifting larvae is vertical positioning (Fiksen et

231

al., in press). It is therefore no surprise that vertical migration of larvae has received a lot of

232

attention, even in early works (Bartsch et al., 1989). Most studies use a diel vertical migration

233

scheme (e.g., Rice et al., 1999; Peliz et al., in press), but other vertical migration approaches

234

include the use of a depth-by-age curve (Ådlandsvik et al., 2004), stage-specific vertical

235

velocities (Pedersen et al., 2006), and length-specific depth distributions (Bartsch and Coombs,

236

2004). Vikebø et al. (in press) have used an approach where larvae are assumed to know the

237

conditions within the upper 100 m and to migrate where they have an optimal blend of growth

238

and survival. Horizontal swimming of larvae has also been considered in a few studies

239

(Rodríguez et al., 2001; Yeung and Lee, 2002; Fox et al., 2006; Fiksen et al., in press). Specific

240

experiments have been conducted to assess swimming abilities of larvae, and described as an

241

integral part of the modelling process (e.g., Guizien et al., 2006). To our knowledge, schooling

9

242

behaviour, which commonly appears in small pelagic fish larvae (Hunter and Coyne, 1982), has

243

never been taken into account. However, there are many models of schooling that focus on how

244

schools form and persist (reviewed in Lett and Mirabet, submitted).

245 246

2.3 Case studies

247 248

2.3.1 Models of anchovy and sardine in the Benguela upwelling system

249 250

Most biophysical models of early life stages developed for the Benguela Current upwelling

251

system have focused on anchovy (Engraulis encrasicolus) in the southern Benguela (Huggett et

252

al., 2003; Mullon et al., 2003; Parada, 2003; Parada et al., 2003; Skogen et al., 2003; Koné,

253

2006). Early life stage models have also been developed for sardine (Sardinops sagax) in the

254

southern Benguela (Miller, 2006; Miller et al., 2006) and in the northern Benguela (Stenevik et

255

al., 2003). All these models, except Koné (2006), fall into the “hydrodynamics and simple

256

behaviors” category; Koné’s (2006) model is an example of an “hydrodynamics and dynamic

257

prey” IBM.

258 259

Most of these Benguela Current models used the same regional PLUME configuration (Penven,

260

2000; Penven et al., 2001) of the ROMS hydrodynamic model (Shchepetkin and McWilliams,

261

2005). The PLUME configuration covers the southern Benguela from 28 to 40°S and from 10 to

262

24°E, at a horizontal resolution ranging from 9 km at the coast to 16 km offshore and with 20

263

terrain-following vertical levels. Koné (2006) used the same PLUME configuration of ROMS but

264

coupled with a biogeochemical model (Koné et al., 2005). Skogen et al. (2003) and Stenevik et

265

al. (2003) used a different configuration and the NORWECOM hydrodynamic model (Skogen,

266

1999). This configuration covered the whole Benguela area, from 12 to 46°S and from 4 to 30°E,

267

at a horizontal resolution of 20 km and with 18 terrain-following vertical levels.

268 269

Recently, the ROMS model has been applied to the Benguela region using an embedding

270

procedure that places a high-resolution small-scale (child) grid nested into a low-resolution large-

271

scale (parent) grid (Penven et al., 2006a). The parent grid covers the whole southern Africa from 10

272

5 to 46°S and from 2°W to 54°E, at a horizontal resolution of 0.25° ranging from 19 km in the

273

south to 27 km in the north (Southern Africa Experiment or SAfE, Penven et al., 2006b). The

274

first child grid (SAfE south coast) was designed to study the interactions between the Agulhas

275

and Benguela systems and covers the area from 28 to 39ºS and 12 to 27ºE at a horizontal

276

resolution of about 8 km (N. Chang, personal communication). The second child grid (SAfE west

277

coast) was designed to encompass most of the Benguela from 18 to 35ºS and 10 to 20ºE, and also

278

has a horizontal resolution of about 8 km (J. Veitch, personal communication). The parent and

279

child grids have 32 terrain-following vertical levels. Hydrodynamics provided on the SAfE west

280

coast child grid was recently used in a Lagrangian model (Lett et al., in press).

281 282

The development of anchovy early life stage models for the southern Benguela that used the

283

PLUME configuration of ROMS has followed a step-by-step progression from simple to more

284

complex models. The first model developed by Huggett et al. (2003) only tracked eggs and larvae

285

with transport completely determined passively by currents. Parada et al. (2003) introduced a

286

buoyancy scheme for the eggs, and Parada (2003) added temperature-dependent and stage-

287

dependent growth and mortality, and vertical swimming behaviour of larvae. A synthesis of these

288

simulation experiments can be found in Mullon et al. (2003). Koné (2006) then added food-

289

dependency to the temperature-dependent larval growth.

290 291

The progression of anchovy model development provides an opportunity for determining how

292

increasing biological complexity affected model results. All of these models used the same

293

configuration of the same hydrodynamic model (PLUME implementation of ROMS), and all

294

relied on the same method for designing simulations and analysing the results. Simulations were

295

performed using all combinations of pre-defined parameter values, and using comparable

296

graphical and statistical analysis to determine the effects of the parameters and their interactions

297

on the output variables. The main output variable used was the percentage of larvae transported

298

from the anchovy spawning grounds near the South African south coast to nursery grounds off

299

the west coast. This percentage was referred to as the simulated transport success. The major

300

results from the progression of models and analyses were:

301 302

1) Simulated transport success has a strong seasonal pattern, peaking in the austral spring and summer time periods (October to March).

11

303 304 305 306 307

2) There is little chance for virtual eggs released on the eastern side of the spawning grounds to be transported to the west coast. 3) Simulated transport success is highest for an egg density of 1.025 g.cm-3, and use of this density generated results similar to those obtained using purely passive transport. 4) Simulated transport success increases with spawning depth within 0 to 75 m.

308 309

These modelling results corresponded reasonably well to the field observations of anchovy in the

310

southern Benguela. The main spawning season of anchovy is austral spring and summer, with

311

major spawning areas located in the western and central parts of the spawning grounds. Measured

312

egg density is about 1.025 g.cm-3. However, anchovy eggs are mainly found in the upper 20 m

313

and rarely deeper than 60 m (Dopolo et al., 2005). Preliminary results obtained from an anchovy

314

biophysical model using new hydrodynamics (SAfE south coast implementation of ROMS)

315

suggest a slight decrease of simulated transport success with spawning depth, which is more in

316

accordance with the observations.

317 318

While the early life stage models of the Benguela described above used hydrodynamics based on

319

seasonal forcing, other models (Skogen et al., 2003; Stenevik et al., 2003; Miller, 2006; Miller et

320

al., 2006) used hydrodynamics based on interannual forcing. Interannual indices of retention or

321

transport success derived from these models did not correlate well with time-series of sardine

322

recruitment (Stenevik et al., 2003; Miller, 2006; Miller et al., 2006). However, the main objective

323

of these analyses was on investigating the factors that affected retention and transport of sardine

324

eggs and larvae, and not on deriving indicators of recruitment success. Skogen et al. (2003)

325

showed that different indices of transport to the South African west coast were correlated to

326

anchovy recruitment. The contrasting results of model-derived indices being correlated or not

327

correlated with recruitment data was also found in other studies. Parada et al. (in press) found that

328

pre-recruitment indices and retention indices derived from their walleye pollock (Theragra

329

chalcogramma) model were not correlated with recruitment estimates in the Gulf of Alaska,

330

while Baumann et al. (2006) found significant correlations between a drift index and sprat

331

(Sprattus sprattus) recruitment success in the Baltic Sea. The recent availability of interannual

332

simulations covering the period 1957–2001 for the SAfE south coast and SAfE west coast

12

333

implementations (N. Chang and J. Veitch, personal communications), presents an opportunity for

334

attempting to correlate model-generated indices to recruitment data for the Benguela region.

335 336

The numerical experiments performed by Mullon et al. (2002), Lett et al. (2006) and Lett et al.

337

(in press) provide some basic building blocks for analyses of anchovy and sardine early life stage

338

dynamics. Mullon et al. (2002) developed an evolutionary model that explored how

339

environmental constraints (e.g., avoiding offshore advection and cold water) affect the spatial and

340

temporal patterns in spawning. These relatively simple constraints imposed for multiple

341

generations led to the selection of spawning patterns that were in surprisingly good agreement

342

with those observed for anchovy and sardine in the southern Benguela. This approach of allowing

343

the temporal and spatial patterns in spawning to emerge from a selective process and

344

environmental conditions offers an alternative to the usual approach of specifying the location

345

and timing of egg deposition that is used in most biophysical models.

346 347

Lett et al. (2006) used a Lagrangian approach to simulate and quantify enrichment and retention

348

processes in the southern Benguela. These two processes, along with the concentration process,

349

form a triad of processes that are fundamental for the survival and recruitment of early life stages

350

of pelagic fish (Bakun, 1996). The results of Lett et al. (2006) reinforce the view of Cape

351

Agulhas as a “dividing line” in the southern Benguela pelagic fish recruitment system, with a

352

transport-based subsystem to the west and a retention-based subsystem to the east. These two

353

subsystems were considered by Miller (2006) and Miller et al. (2006) in their sardine biophysical

354

model. Lett et al. (in press) also used a Lagrangian approach to investigate the processes

355

responsible for the absence of anchovy and sardine spawning in the central Benguela off the

356

Lüderitz region. They examined the flow field and temperature conditions that particles would

357

experience in the Lüderitz region, and concluded that the combination of a surface hydrodynamic

358

and a subsurface thermal barrier could limit the possibility for anchovy and sardine

359

ichthyoplankton to be transported from the southern to the northern Benguela. Recent remote

360

sensing data also suggested a poor trophic environment in the Lüderitz region (Bartholmae and

361

Basson, submitted; Demarcq et al., in press).

362

13

363

2.3.2 Models of anchovy and sardine in other upwelling systems

364 365

We are not aware of biophysical models of anchovy early life stages developed in Eastern

366

Boundary Current upwelling systems outside the Benguela. A preliminary biophysical model of

367

sardine (Sardina pilchardus) has been developed in the Iberian system (Santos et al. 2005). Other

368

models of sardine, in the Kuroshio Current (Sardinops melanostictus, Heath et al., 1998) and in

369

the Iberian (Santos et al., 2004) upwelling systems, used hydrodynamics derived from field data.

370

However, three anchovy (Engraulis ringens) biophysical models using hydrodynamics provided

371

by ROMS are under development for the Humboldt Current upwelling system (Brochier et al.,

372

submitted; Brochier et al., abstract 2 ; Soto-Mendoza et al., abstract 3 ; Chai et al., abstract 4 ), and

373

there are anchovy models in the vicinity of major upwelling systems. Models exist for European

374

anchovy (Engraulis encrasicolus) in the Bay of Biscay (Allain et al., 2003; Allain, 2004; Allain

375

et al., in press; A. Urtizberea, personal communication) and for Japanese anchovy (Engraulis

376

japonicus) in the Yellow Sea (Hao et al., 2003). Models of the early life stages of taxa in

377

upwelling systems besides small pelagic fish include crab (Carcinus maenas) in the Iberian

378

upwelling system (Marta-Almeida et al., 2006; Peliz et al., in press), zooplankton in the

379

California Current upwelling system (Carr et al., in press), and blue shrimp (Litopenaeus

380

stylirostris) and brown shrimp (Farfantepenaeus californiensis) in the Gulf of California

381

(Marinone et al., 2004).

382 383

Santos et al. (2004) used a combination of measured and modelled velocities to estimate the

384

surface flow field experienced by particles during an upwelling event off Portugal, and obtained

385

qualitatively similar patterns of retention at the shelf-break for modelled particles and for

386

observed sardine eggs and larvae. In the same region, Marta-Almeida et al. (2006) used the 2

Brochier, T., Tam, J. and Ayón, P. IBM for the anchovy in the northern Humboldt Current ecosystem: identification of processes affecting survival of early life stages. International Conference on The Humboldt Current System: Climate, ocean dynamics, ecosystem processes, and fisheries. Lima, Peru, November 27 - December 1, 2006. 3

Soto-Mendoza, S., Castro, L., Parada, C., Colas, F., Donoso, D. and Schneider, W. Modeling the egg and early larval anchoveta (Engraulis ringens) transport/retention in the southern spawning area of the Humboldt Current. International Conference on The Humboldt Current System: Climate, ocean dynamics, ecosystem processes, and fisheries. Lima, Peru, November 27 - December 1, 2006.

4

Chai, F., Shi, L., Xu, Y., Chao, Y., Rose, K., Chavez, F. and Barber, R. T. Modeling Peru upwelling ecosystem dynamics: from physics to anchovy. International Conference on The Humboldt Current System: Climate, ocean dynamics, ecosystem processes, and fisheries. Lima, Peru, November 27 - December 1, 2006.

14

387

ROMS model on a domain extending from 37.5 to 44.3°N and 12.8 to 8.4°W and simulated

388

different vertical migration behaviours of crab larvae. They showed that migrating larvae were

389

retained within the inner shelf under a wider range of upwelling and downwelling conditions than

390

non-migrating larvae. This was further studied by Peliz et al. (in press). They used a 3-level

391

embedding procedure in ROMS, with a high-resolution (~2 km) small-scale domain nested into a

392

medium-resolution (~6 km) medium-scale domain, both nested into a low-resolution (~16 km)

393

large-scale domain. They showed that a model including both diel vertical migration for larvae

394

and river outflow allowed simulating patterns of crab larvae concentrations that were closer to the

395

observed patterns than when only one of the two processes was included (Figure 2). This is a

396

good example of modelling study using a pattern-oriented approach (Grimm et al. 2005). We can

397

anticipate a similar study on sardine (Santos et al., 2007) to complement the preliminary one of

398

Santos et al. (2005), as data on the vertical distribution and behaviour of sardine larvae have

399

recently been collected off Portugal (Santos et al., 2006).

400 401

Figure 2

402 403

Allain and colleagues (Allain et al., 2003; Allain 2004; Allain et al., in press) used

404

hydrodynamics derived from the MARS 3-dimensional circulation model (Lazure and Dumas, in

405

press) that covered the Bay of Biscay south of 49°N and east of 8°W at a 5 km horizontal

406

resolution with 30 terrain-following vertical levels. Allain et al. (2003) used virtual passive buoys

407

released in the simulated flow fields to reconstruct the putative origins in time and space of

408

collected anchovy larvae and juveniles. Allain (2004) and Allain et al. (in press) used a

409

biophysical model of anchovy that included transport, growth, and mortality. They used the

410

super-individual approach (Scheffer et al., 1995), with each simulated buoy representing a large

411

number of eggs, larvae, and juveniles. Every time step, each buoy was characterized by a

412

distribution of growth rates (function of age and the physical environment it experienced) and by

413

a survival probability (the probability that growth rates in the distribution were above a pre-

414

defined age-specific and stage-specific threshold). They used the simulations to form an index of

415

recruitment (survival rate multiplied by egg production, summed over the year and spatial

416

domain) that correlated remarkably well to a (short) time-series of anchovy recruitment data.

417

15

418

Rodríguez et al. (2001) released particles in a static two-dimensional flow field calculated from

419

observations to assess the potential for retention of organisms around the Canary Island of Gran

420

Canaria. They showed that particles released north of Gran Canaria would need limited

421

“swimming ability” (speeds of ~0.5 cm.s-1) in order for them to be retained, whereas those

422

particles released east or west of Gran Canaria would require a much stronger swimming ability

423

(>5 cm.s-1) in order to be retained. Hunter (1977) reported mean swimming speeds of 0.3–0.45

424

cm.s-1 for early (5 mm) northern anchovy (Engraulis ringens) larvae at 17° to 18°C. A

425

configuration of ROMS for the northern Canary is under development (E. Machu, personal

426

communication). This configuration uses an embedding procedure with a parent grid covering the

427

area from 10 to 40°N and -5.5 to 30°W at a horizontal resolution of 1/4°, and a child grid from 21

428

to 32.5°N and 9 to 20.5°W at a horizontal resolution of about 8 km. A biophysical model will be

429

developed in the near future to study sardine early life history in this region (T. Brochier and A.

430

Ramzi, personal communications).

431 432

Lett et al. (2007) used a similar approach for the northern Humboldt upwelling system as Lett et

433

al. (2006) did in the southern Benguela. They derived indices of enrichment, concentration, and

434

retention based on Lagrangian simulations using hydrodynamics from a regional ROMS

435

configuration. This configuration covered the domain off Peru from 20°S to 3°N and 90 to 70°W

436

at a horizontal resolution of 1/9° (~10 km) and with 32 terrain-following vertical levels (Penven

437

et al., 2005). Lett et al. (2007) analysed the spatial distribution and seasonal variability in the

438

simulated indices, and discussed their results in relation to the distributions of anchovy

439

(Engraulis ringens) eggs and larvae off Peru. A coastal area of enhanced enrichment was

440

identified between Punta Falsa and Pisco (6°–14°S), which corresponds to the zone where most

441

anchovy eggs and early larvae are found. Preliminary modelling results suggest that the early life

442

stages of anchovy are found mainly in the northern part of this zone (6°–9°S) because this area

443

provides high retention and concentration. Another striking characteristic of anchovy spawning in

444

the region is the bimodal seasonal pattern, with peaks in January-March and August-October. A

445

biophysical model of anchovy early life stages is currently being developed to better understand

446

the bimodal spawning pattern in this region (Brochier et al., submitted). Other biophysical

447

models of the early life stages of anchovy are under development for the northern Humboldt (Y.

16

448

Xu, personal communication) and southern Humboldt (S. Soto-Mendoza, personal

449

communication) upwelling systems.

450 451

Heath et al. (1998) used static two- and three-dimensional flow fields calculated from

452

observations for two years to assess the dispersal of sardine (Sardinops melanostictus) eggs and

453

larvae in the Kuroshio Current upwelling system. They obtained very different results for the two

454

years, with most egg production being transported towards the Kuroshio Current extension in the

455

first year, while most of it was retained in the coastal areas in the second year. In contrast, using

456

different scenarios for transport (without and with diffusion) and for behaviour (without and with

457

vertical redistribution of individuals according to observed depth distributions) had little effect on

458

model simulations. They argued that, at least during the time of their study, hydrodynamic

459

models could not easily simulate the instabilities in the Kuroshio path that were responsible for

460

the contrasting flow fields they observed.

461 462

Hao et al. (2003) conducted Lagrangian simulations in the nearby Yellow Sea using a regional

463

configuration of the HAMSOM (Hamburg Shelf Ocean Model) hydrodynamic model. This

464

configuration covered the Yellow Sea and the East China Sea from about 25 to 40°N and 120 to

465

130°W at a horizontal resolution of 1/12° and with 12 vertical layers. They showed that including

466

a tidal component in the hydrodynamic model affected the predicted transport pattern of

467

simulated particles, and discussed their results in relation to the high concentrations of anchovy

468

(Engraulis japonicus) eggs and larvae that are observed in tidal fronts.

469 470

Despite the preliminary Lagrangian experiments conducted in the California Current upwelling

471

system to investigate the effects of vertical migration on larval drift trajectories (Botsford et al.,

472

1994), there has not been much follow-up with more biologically-based models of the early life

473

stages of small pelagic species. Lagrangian experiments have been designed to study the transport

474

patterns of other taxa, including zooplankton (Carr et al., in press) and shrimp larvae (Marinone

475

et al., 2004). Carr et al. (in press) used a nested approach in ROMS with a parent grid covering

476

the entire California upwelling system (7.5 km resolution) and a child grid focusing on the

477

Monterey Bay region (2.5 km resolution), and with 32 terrain-following vertical levels. They

478

investigated the effects of diel vertical migration of planktonic organisms, and obtained results

17

479

suggesting that migration into subsurface onshore currents during the day would not compensate

480

for surface offshore transport during the night and thus retention would be low within the

481

Monterey Bay. Marinone et al. (2004) explored different scenarios for when advection of

482

particles occurred (only during the day, the night, or when the current flows northwards), and

483

discussed their results in relation to transport patterns of shrimp larvae to their nursery areas.

484

They used a configuration of HAMSOM, developed by Marinone (2003), that covered the Gulf

485

of California at a horizontal resolution of 1/24° (~4 km) and with 12 vertical layers.

486 487

2.3.3 Models of small pelagic fish in “non-SPACC” regions

488 489

In this section, we briefly mention several early life stage models for small pelagic species and

490

regions outside the main focus of SPACC. We mention these examples because the methods and

491

results may be of interest, and because they provide a broad view of early life stage biophysical

492

models.

493 494

Vaz et al. (submitted) used outputs from a western South Atlantic configuration of the Princeton

495

Ocean Model (POM) as input to a biophysical model of anchovy (Engraulis anchoita) eggs and

496

early larvae. They investigated the spatial and seasonal patterns in the levels of retention and

497

transport along the coast off northern Argentina, Uruguay, and southern Brazil.

498 499

Bartsch and colleagues (Bartsch et al., 1989; Bartsch, 1993; Bartsch and Knust, 1994a, 1994b)

500

used hydrodynamics derived from early versions of the HAMSOM model as input to biophysical

501

models of herring (Clupea harengus) and sprat (Sprattus sprattus) early life stages for the North

502

Sea. Their seminal studies incorporated transport and size-dependent vertical migration of larvae,

503

and focused on qualitative comparisons between observed and simulated distributions of larvae.

504

Sætre et al. (2002) used the NORWECOM model to assess the transport routes and retention

505

areas for herring along the Norwegian coast. They showed that herring tended to spawn in areas

506

where retention was enhanced by the presence of topographically-induced quasi-stationary

507

eddies. Hinrichsen et al. (2005) developed a model of sprat for the Baltic Sea and showed that the

508

degree of overlap between observed and simulated distributions of juveniles was higher when

18

509

individuals were constrained to remain close to the surface than when they were allowed to

510

migrate to deeper waters during the day. Baumann et al. (2006) proposed different empirical

511

models fitted to observed values of sprat recruitment, and obtained the best result when they

512

incorporated a drift index derived from the biophysical model simulations.

513 514

Bartsch and Coombs (2001) developed a mackerel (Scomber scombrus) early life stage

515

biophysical model using an eastern North Atlantic configuration of the HAMSOM model. They

516

used initial conditions for egg distribution and abundance that were based on field data, and

517

simulated the drift of eggs and larvae and their growth using a function dependent on temperature

518

and age. Bartsch and Coombs (2004) later introduced additional biological processes into their

519

model, including vertical migration, feeding, and mortality. They used size-dependent diel

520

vertical migration, and dynamic prey fields calculated from satellite-derived sea-surface

521

temperature and chlorophyll-a concentration. Like Allain (2004), Bartsch and Coombs (2004)

522

also used super-individuals, where every simulated entity initially represented 106 individuals

523

who experienced the same environment and who died according to growth-dependent and length-

524

dependent mortality rates. Using this model, Bartsch et al. (2004) derived indices of mackerel

525

early post-larvae survival that they then compared to juvenile catch data for different sub-areas

526

during 1998 to 2000. These simulations however all used the same initial conditions of egg

527

distribution and abundance (those observed for 1998). Bartsch (2005) showed that simulating

528

survival for 2001 using 1998 egg data as the initial conditions lead to a “wrong” 60% increase in

529

survival compared with using actual 2001 egg data.

530 531

3 Biophysical models that include adults

532 533

There are two major categories of models that include adult fish and physics. The first category is

534

models in which the key processes of growth, mortality, reproduction, or movement depends on

535

the physics or variables greatly influenced by the physics. The second category is models for

536

which the questions to be addressed require full life cycle simulations (i.e., include all life stages)

537

so that the long-term (multigenerational) effects of impacts or environmental changes can be

538

examined. These full life cycle models thus must deal with eggs, larvae, and adults in a single

19

539

model. Such models are relatively rare now but model development is headed in that direction.

540

We expect biophysical full life cycle models to be the focus of much effort during the next

541

decade, and anticipate that full life cycle biophysical models will be especially important for

542

addressing issues related to the effects of climate change on fish populations and fisheries.

543 544

In this section, we review biophysical models that include adult stages of fish. We focus our

545

review on those models that are spatially-explicit, and that use physics, or variables derived from

546

the physics, as inputs. We also include the NEMURO family of models because some versions of

547

the NEMURO models fit our criteria for inclusion, but also because this effort is ongoing and

548

provides an example of a spatially-explicit full life cycle approach. We do not include models

549

that use environmental variables as inputs (e.g., temperature) but only simulate a single spatial

550

box (e.g., Robinson and Ware, 1999; Clark et al., 2003), and also do not include other population

551

modelling approaches (e.g., spawner-recruit models, Fiksen and Slotte, 2002; surplus production

552

models, Jacobson et al., 2005). These other modelling approaches are likely useful for addressing

553

certain questions related to climate change effects on fish. Our focus in this chapter is on

554

biophysical spatially-explicit models that include adult fish.

555 556

3.1 Abiotic and biotic environment, growth, and mortality

557 558

Models that include adults generally use similar information from the physics and lower trophic

559

models as used by the egg and larval models. These outputs include: current velocities

560

(advection), salinity, temperature, and prey fields. The growth and mortality formulations in adult

561

models are also similar to those by the early life stage models. Growth in weight is generally

562

either based on empirical relationships (e.g., von Bertalanffy equation, Shin and Curry, 2004) or

563

bioenergetics-based (e.g., Megrey et al., 2007), and natural mortality rate is often treated as a

564

constant. There are also some predator-prey biophysical models that include adults wherein

565

mortality depends on fish encounters with other fish (Shin and Cury, 2004; Ault et al., 1999).

566

With the inclusion of adult life stages, representing harvest rates and fisheries can also become

567

important.

568

20

569

3.2 Movement and behaviour

570 571

The process of movement is where models that include adults differ from early life stage models.

572

Unlike the egg and larval stages, movement of juvenile and adult fish no longer is necessarily

573

dominated by the physics. Movement of adults can be influenced by the physics, neutral with

574

respect to the physics, or even opposite to the physics. In many situations, the physics (or

575

variables directly derived from the physics, e.g., salinity, temperature, food) influences some

576

portion of the fish’s movement, while another component is due to other factors dependent on

577

behavioural decisions. An extreme example of movement of adult fish opposite to the physics is

578

migration upstream or counter to the prevailing currents.

579 580

How to represent the movement of adult fish in spatially-explicit biophysical models is unclear

581

and there are several approaches that have been proposed. Conceptually, movement of adults

582

adds terms to equation 1 attributable to behaviour, and then there must be some weighting of the

583

relative contribution of the physics-based term and the behaviour-based term. We highlight here

584

two of the approaches for modelling movement to illustrate some of the variety in how one can

585

represent movement of adults in biophysical models.

586 587

Railsback et al. (1999) proposed an approach based on fitness considerations in which individuals

588

search neighbouring cells, and move towards cells that provide an optimal blend of growth and

589

survival. In our context, the physics would provide the temperature and prey fields, and possibly

590

other variables, that are needed to compute growth and mortality in the neighbouring cells. The

591

growth and mortality in each neighbouring cell is then combined into a fitness measure that is the

592

projection of the likelihood of an individual surviving and obtaining some target size at some

593

time in the future (e.g., size at maturity next year), assuming conditions in the cell remained

594

constant into the future. Individuals move towards the cell in their neighbourhood that has the

595

highest fitness.

596 597

Humston et al. (2004) implemented a different approach to modelling movement of fish based on

598

the concept of kinesis that does not require that the fish has knowledge of the conditions in

599

neighbouring cells. Rather, individuals evaluate the conditions in their present cell against some

21

600

specified optimal value. Movement has inertial and random components; inertial movement

601

involves smaller angles and shorter distances than the random component. They used a weighting

602

scheme to combine the inertial and random components into distances moved and the angle of

603

movement. The closer the conditions in the present cell were to optimal, the more the inertial

604

movement dominated and fish would make smaller moves going in about the same direction and

605

therefore tended to stay in the good areas. Poor conditions resulted in the random movement

606

being weighted more heavily and the appearance of individuals searching over relatively large

607

areas and in all directions.

608 609

The Railsback and Humston approaches illustrate a fundamental schism in movement modelling

610

approaches: whether organisms can sense the conditions in neighbouring cells sufficiently to

611

project how conditions there would affect their growth or mortality (Tyler and Rose, 1994). With

612

an appropriate weighting scheme, one can also add terms for the contribution of advective

613

transport to the mix in both approaches. Other approaches to modelling movement include multi-

614

step approaches that first determine the type of behaviour from a list of discrete options and then

615

implement the specifics of movement associated with that behaviour (Blackwell, 1997; Anderson,

616

2002; Goodwin et al., 2006), and the use of genetic algorithms to train neural networks (Huse and

617

Giske, 1998).

618 619

Given the uncertainty in how to represent movement, it is important to note that some biophysical

620

models bypassed the issue. Heath et al. (1997) did not try to model movement but rather simply

621

assumed movement would occur and would result in the spatio-temporal distributions derived

622

from field data. Even more extreme is the growth potential approach (Luo et al., 2001). They

623

ignored movement and spatial distributions, and simply predicted what would happen if fish were

624

present in all locations.

625 626

3.3 Case studies

627 628

We have organized this section differently from the early life stage case studies section. There are

629

far fewer examples that include physics and adult fish, and not enough examples to divide by

22

630

species and geographic area as was done with the early life stage examples. Rather, we have

631

grouped the adult models together according to two general categories: models of adults that use

632

physics and full life cycle models. We emphasize the methods used because of the much wider

633

diversity of approaches used with adult models than with the more standardized approaches

634

generally used with early life stage models.

635 636

3.3.1 Models of adults that use physics or physics-related variables

637 638

Luo et al. (2001) used the output of a 3-dimensional hydrodynamic model coupled to a water

639

quality model as input to a fish feeding and bioenergetics model to determine the carrying

640

capacity of Chesapeake Bay for young-of-the-year juvenile menhaden (Brevoortia tyrannus). The

641

water quality model had over 4000 cells, arranged as a surface grid of 729 cells, with vertical

642

cells every 2 m. They averaged the 2-hour simulated values of temperature, dissolved oxygen,

643

and chlorophyll-a to obtain daily values for each cell for June through December in an average

644

freshwater inflow year. Temperature was used directly in the bioenergetics model, and affected

645

maximum consumption rate and respiration rate. Menhaden are filter feeders and so chlorophyll-a

646

concentrations were multiplied by the area of the mouth, swimming speed, and an efficiency term

647

to obtain realized consumption rate. Consumption, with the rest of the bioenergetics model,

648

enabled prediction of daily growth rate in each cell. Carrying capacity for each cell was then

649

computed based on the predicted consumption rate, prey production rates, and prey biomass, and

650

further adjusted for low dissolved oxygen, to obtain the biomass of menhaden that could be

651

supported in that cell on each day.

652 653

Luo et al. (2001) showed spatial maps and reported the percent of the bay volume able to support

654

different growth rates and carrying capacities. Their results showed that there was large spatial

655

and temporal variation in growth rate potential and biomass supportable due to the nonlinear

656

functional forms in the feeding and bioenergetics models and as a result of combining the effects

657

of the multiple factors of temperature, chlorophyll-a, and dissolved oxygen. Menhaden must

658

occupy cells with growth potential greater than 0.005 to 0.001 g/g/day during the June through

23

659

December growing season in order to achieve their observed weights in December, and the daily

660

total volume of such good habitat in the bay varied between practically zero and 80%.

661 662

There are many examples of the use of habitat suitability indices (HSI) that use spatially-explicit

663

environmental variables estimated from field data or outputted from hydrodynamic models. Most

664

of these examples involve how changes in stream and river flows would affect the habitat for

665

specific species downstream because the HSI approach was initially developed for evaluating

666

how water releases from dams would affect fish downstream (Acreman and Dunbar, 2004). Use

667

of HSI avoids the issues of representing how adult fish move because an endpoint can be simply

668

the changes in the quantity and quality of the habitat. The HSI approach has also been used as an

669

intermediate variable that is then used to affect movement of adult fish (e.g., Lehodey et al.,

670

2003).

671 672

Rubec et al. (2001) offers an example of HSI applied in an estuarine environment. We summarize

673

an estuarine example here even though it uses field data to derive the environmental variables

674

because the same analysis would be used with model-predicted environmental variables. Others

675

have computed HSI values from the output of hydrodynamic and water quality models (e.g.,

676

Guay et al., 2000). Rubec et al. interpolated temperature and salinity values to obtain seasonal

677

values for 18.5 m2 cells in the surface and bottom layers for Tampa Bay, Florida. Additional

678

information on depth, bottom substrate type, and species abundances were also used.

679

Relationships between suitability (0 to 1) and each variable were formulated from abundance

680

data, and the product of the suitabilities in each cell (geometric mean) was computed as the

681

overall suitability of that cell for that life stage. They then checked whether overall suitability was

682

correlated to species abundances over the four seasons, and swapped suitability functions with

683

those developed for Charlotte Harbour to see if suitability functions were transferable among

684

locations.

685 686

Karim et al. (2003) coupled a hydrodynamic-NPZ model to a model of fish movement and

687

dissolved oxygen-related mortality. Their simulations were based on the Marbled Sale

688

(Pleuronectes yokohamae), a demersal species, in Hakata Bay. The spatial domain was a

689

horizontal grid of 300 m x 300 m cells with five vertical layers. The grid was selected to

24

690

correspond to the horizontal distance typically traveled by an individual fish in the 30 minute

691

time step. Movement could then be based on individuals moving to neighboring cells each time

692

step. The hydrodynamic-NPZ model was solved using its own time step, and temperature and

693

dissolved oxygen values were obtained for each grid cell for use with the fish movement and

694

mortality models. Karim et al. used a series of laboratory experiments and field tracking data of

695

individual fish to develop and calibrate the movement model. Movement in the model was based

696

on computing the preference of each neighboring cell in 3 dimensions from functions that related

697

temperature to a preference level and dissolved oxygen concentration to a preference level, and

698

then a function that combined the preference levels into a single overall preference value for the

699

cell. Individuals moved to the cells with the highest overall preference. This approach to

700

modeling movement could be considered a simplified version of the more general Railsback et al.

701

(1999) fitness-based approach. Horizontal position was updated every 30 minutes, while vertical

702

position was evaluated every 10 seconds. Exposure to low dissolved oxygen caused increased

703

mortality.

704 705

After performing simulations that satisfactorily mimicked the laboratory and field tracking

706

results, Karim et al. (2003) performed 3-day simulations with individuals released in different

707

vertical layers (surface or bottom) and horizontal locations (whole grid or inner bay). They

708

computed the mortality rate of the cohort from exposure to low dissolved oxygen as mortality

709

accumulated over the three days. They concluded that hypoxic conditions in the inner portion of

710

the bay during the summer can cause significant mortality of demersal fish, and that future model

711

developments should include the simulation of hypoxia effects on pelagic species.

712 713

Ault et al. (1999) developed a predator-prey model of spotted seatrout (Cynoscion nebulosus) and

714

pink shrimp (Penaeus duorarum) that was imbedded in a 2-dimensional hydrodynamic model of

715

Biscayne Bay, Florida. The hydrodynamic model consisted of 6346 triangular elements and 3407

716

nodes, with grid spacing between nodes of about 500 m. The model was driven with tide, winds,

717

and freshwater discharge data for 1995. A 1 minute time step was used to solve the hydrodynamic

718

model, and currents and salinities were output at 10 minute intervals for use with the fish model.

719

Shrimp and seatrout were followed as super-individuals (what Ault et al. termed “patches”).

720

Patches were introduced monthly for shrimp, and at night on incoming tides for seatrout. Patches

25

721

were tracked using continuous x and y positions; each time step the patch experienced the

722

conditions of the cell it inhabited. Circulation was used to move around the patches of egg and

723

yolk-sac larval seatrout, circulation and behavior together was used with pre-settlement shrimp,

724

and active behavior only was used for moving settled shrimp and juvenile and adult seatrout.

725

Salinity affected the angle and distance moved (based on swimming speed) of pre-settled shrimp,

726

which was added to the circulation-based movement. Horizontal movement only occurred at

727

night; shrimp went to bottom during the day. Once shrimp settled to the bottom, they were moved

728

around based on their body length and habitat quality. Seatrout movement was based on the

729

growth rate in cells within a specified distance (detection range) of their present location, and

730

patches moved towards cells with the highest growth potential. Movement within the Ault et al.

731

model of the different life stages of shrimp and seatrout illustrates the diversity of movement

732

algorithm and approaches, including purely hydrodynamics-driven, an approach rooted in cellular

733

automation, an approach that was later generalized into the Humston et al. (2004) kinesis model,

734

and a hybrid mix of the kinesis and Railsback’s fitness-based approach. Shrimp grew based on

735

the cumulative temperature they were exposed to; seatrout grew based on bioenergetics with

736

consumption based on ingested shrimp co-located in their present cell.

737 738

Ault et al. (1999) reported the results of one year simulations that illustrated model behavior and

739

the usefulness of the model for examining how environmental and biological factors affect

740

predator-prey dynamics. They concluded that June-spawned seatrout were transported more into

741

the bay, settled over a wider range of habitats, and grew faster than August-spawned cohorts.

742

They also showed that spatial variation in habitat quality can affect growth rates of fish, even

743

causing difference among individuals from within the same spawning cohort.

744 745

3.3.2 Full life cycle models

746 747

The EcoPath with EcoSim (EwE) family of models (Walters et al., 1997) are biomass-based

748

compartment models and consist of a steady-state version (EcoPath), which is also used as initial

749

conditions for a single spatial box time-dynamic version (EcoSim). Walters et al. (1999) extended

750

the EwE models to be spatially-explicit (EcoSpace) by imbedding an EcoSim model on each cell

26

751

of a 2-dimensional grid of cells. Differential equations are constructed for each fish compartment

752

(species or functional group) in an analogous manner as the equations for phytoplankton and

753

zooplankton in NPZ models (net effect of gain and loss terms corresponding to growth and

754

mortality processes). In early versions of EwE, fish were represented as biomass, which was

755

recognized as inadequate because of the large ontogenetic changes in growth and mortality in

756

many fish species and because a single biomass compartment did not allow for explicit

757

representation of recruitment dynamics which is critical to understanding and forecasting fish

758

population dynamics. Thus, EcoSim was extended to allow for each fish compartment to be

759

subdivided into smaller compartments (e.g., juveniles versus adults) as a crude way to allow for

760

ontogenetic differences and to make recruitment dynamics semi-distinct from total biomass

761

(Walters et al., 2000). Recent versions of EcoSpace went further and allowed for monthly cohorts

762

(age-structure) of key species or groups. EcoSpace is presently undergoing modification to allow

763

for even more detailed representation (e.g., following individuals or packets) for selected taxa,

764

and for easy coupling to hydrodynamic and NPZ models. While there have been many

765

applications of EcoPath and quite a few applications of EcoSim (e.g., Shannon et al., 2004; Field

766

et al., 2006), the spatially-explicit EcoSpace version has not yet received much attention. We

767

mention EwE here because we anticipate increasing attention being paid to the EcoSpace

768

approach in the future.

769 770

One example of the use of the EwE models is that of Shannon et al. (2004), who applied EcoSim

771

to the Southern Benguela ecosystem. The model was comprised of about 30 species or groups,

772

and they adjusted previous parameter values and assumptions (e.g., see Shannon et al., 2003)

773

based on the model’s ability to replay the historical biomasses of selected species for the 1978 to

774

2002 period. Using the EwE fitting algorithm, Shannon et al. (2004) explored how changes to

775

historical fishing pressure patterns, prey vulnerabilities to predation of key prey-predator

776

combinations, and environmental forcing of primary production affected the model fit to the

777

observed biomass time series. The fit of modelled biomasses to the observed time series was

778

insensitive to alterations in the fishing pressure time series, and moderately sensitive to

779

interannual variation in environmental forcing of primary production. The model’s relatively high

780

sensitivity to prey-predator vulnerabilities was consistent with the idea of wasp-waist control

27

781

(Cury et al., 2000), in which small pelagic fish control their zooplankton prey while

782

simultaneously exerting bottom-up control on their predators.

783 784

The IGBEM and BM2 models (Fulton et al., 2004a, 2004b) are also biomass-based but they are

785

truly spatially-explicit, coupled to an elaborate water quality model, and separate each fish

786

species or group into age-classes and follow the average weight and numbers of individuals in

787

each age-class. BM2 was developed in an attempt to reduce the complexity and parameter

788

demands of its more complicated predecessor IGBEM (Fulton et al., 2004b). The application to

789

Port Phillip Bay, Australia, followed 29 living compartments, plus compartments related to

790

detritus, nutrients, sediment, and some physical variables in a 3-layer grid with about 59 cells in

791

each layer. Transport among cells was derived from the output of a hydrodynamic model. A daily

792

time step was used, although if rates were too fast, then a finer time step was adopted. Four fish

793

groups were followed, with each represented by means of an age-structured cohort approach.

794

Spawning occurred outside of the model domain and recruits were injected into the grid on a

795

specific day as the initial number of individuals in the first age-class and then the oldest age-class

796

was removed. The average body weight of an individual in each age-class was followed using

797

two separate weight variables (structural and reserve) based on simulated growth. Growth

798

depended on the summed prey biomass, whose vulnerability was sized-based, in the same cell as

799

the predator and adjusted for feeding efficiency and crowding. Consumed prey was imposed as

800

mortality on the appropriate prey compartments in the cell. Movement of fish age-classes was

801

simulated using fluxes among cells based on specified quarterly target densities in cells and how

802

many days were left in the quarter. Fulton et al. (2004b) acknowledged that the recruitment and

803

movement formulations of the model were likely the weakest aspects of their model, and they

804

have investigated using spawner-recruit relationships and forage-based and density-based

805

movement approaches. They stated that they used the constant recruitment and simple movement

806

because the results did not differ much between the simple and more complicated alternatives.

807 808

Fulton et al. (2004b) performed a variety of simulations of BM2 and compared the outputs to

809

field data for Prince Philip Bay, to field data for other estuaries, and to predictions from the more

810

complicated IGBEM version. Simulations repeated a 4-year time series of forcings as input and

811

they determined that 30-year simulations were sufficient because the model state after 30 years

28

812

was similar to that predicted after 100 years. Examples of model outputs compared to field data

813

or to IGBEM predictions included: averaged biomasses of key compartments, predicted

814

community composition, relationship between DIN and chlorophyll-a, predicted and expected

815

size-spectra features, system-wide indices such as P/B ratios and cycling indices, and the

816

temporal and spatial dynamics of key compartments.

817 818

Adamack (2007) and Adamack et al. (abstract 5 ) coupled a 3-dimensional hydrodynamic-water

819

quality model to an individual-based population model of bay anchovy (Anchoa mitchilli) in

820

Chesapeake Bay. The water quality model is a 3-dimensional model that simulated 24 state

821

variables, including dissolved oxygen, four forms of nitrogen, four forms of phosphorous, two

822

phytoplankton groups, and two zooplankton groups in a 4,073 (729 surface layer cells) cell grid.

823

Anchovy were introduced weekly during the summer spawning season as individual recruiting

824

juveniles and followed until they reached their end of third year when they were removed from

825

the model. Growth and mortality rates of individual bay anchovy within a water quality model

826

cell were calculated every 15 minutes. Growth depended on a bioenergetics model with the

827

predicted zooplankton densities from the water quality model providing the prey for bay anchovy

828

consumption. Anchovy consumption was summed by cell and, with the predicted diet, imposed

829

as an additional mortality term back onto the zooplankton. Anchovy excretion and egestion also

830

contributed to the nutrient recycling dynamics of the water quality model. Mortality rate of

831

anchovy was assumed to decrease with their length. Movement of individual anchovy was

832

simulated both vertically (hourly) and horizontally (daily) based upon temperature, salinity, and

833

prey densities using the kinesis approach of Humston et al. (2004).

834 835

Adamack (2007) reported the results of simulations that used the dynamically coupled models to

836

predict the effects of changes in nitrogen and phosphorous loadings on bay anchovy growth rates

837

and survival. Ten-year simulations using historical sequences of low, average, and high

838

freshwater inflow years were performed under baseline, increased, and reduced nutrient loadings.

839

Results showed that the anchovy response to changes in nutrient loadings was a complex function

5

Adamack, A. T., Rose, K. A. and Cerco, C. F. Simulating the effects of nutrient loadings on bay anchovy population dynamics in Chesapeake Bay using coupled water quality and individual-based models. ECSA 41st International Conference - Measuring and Managing Changes in Estuaries and Lagoons, Venice, Italy, October 1520, 2006.

29

840

of changes in high-quality habitat, prey densities, assumptions about movement, and the

841

magnitude and temporal pattern of the introduction of young-of-the-year recruits. This analysis

842

provides an example of a biophysical model in which the physics is not used directly because the

843

egg and larval stages were bypassed and the adults did not need circulation information, but the

844

physics was needed to properly simulate the NPZ portion.

845 846

Lehodey et al. (2003) used the output of a 3-dimensional hydrodynamic and NPZ model of the

847

Pacific Ocean (45-65oN; 100oE-70oW) as input to an Eulerian-based tuna population dynamics

848

model (SEPODYM). The NPZ model used cells that were 2o in longitude by 2o in latitude at the

849

extreme north and south boundaries of the model domain and 0.5o square near the equator. Forty

850

vertical layers were represented with a layer every 10 meters within euphotic zone and thicker

851

below. Predicted currents were averaged over the 0-30 m surface layer, primary production was

852

integrated over the euphotic zone (1 to 120 m) and with SST, were interpolated on 2-dimensional

853

grid of 1-degree resolution. Lehodey et al. added separate population models for tuna and for

854

what they termed “tuna forage”. The forage model simulated the biomass of tuna prey in each cell

855

using advection and diffusion, and assuming continuous recruitment based on new primary

856

production predicted by the NPZ model and a time lag to account for the delay until the primary

857

production would show up as new forage biomass. The mortality rate and time lag of the forage

858

were related to SST. Tuna population dynamics were modelled by following the numbers in age-

859

classes. Length-at-age were determined from a von Bertalanffy relationship; length was then

860

converted to weight. Two habitat indices (adult and spawning) were computed for each cell based

861

on SST in the cell. The spawning index was used to divide up the annual recruitment among the

862

individual cells. For larvae and juveniles (i.e., until about 4 months of age), tuna movement was

863

purely advection-diffusion. Movement of adult tuna also used the advection terms but that were

864

adjusted by tuna length and by the adult habitat index. They assumed movement was proportional

865

to the length of the age-class and increased with poor habitat quality in the local cell. Natural and

866

fishing mortality was applied to each age-class, with natural mortality rate increased when the

867

habitat was poor (spawning index used for first age-class; adult index used for the rest of the

868

ages). Multiple (six) fisheries with specific gear types and with effort varying by month and by

869

cell or subregion were simulated to drive the fishing mortality rate. The SEAPOPDYM

870

application shared information and used outputs from a traditional fisheries stock assessment

30

871

model called MULTIFAN-CL (Fournier et al., 1998). Annual recruitment in SEAPOP was

872

calibrated to match the overall recruitment estimated by MULTIFAN-CL.

873 874

Lehodey et al. (2003) reported the results of a NPZ and SEPODYM simulation that spanned 1960

875

to1999. General features of the simulation were described, such as the effects of ENSO events

876

and the 1976-77 regime shift, to check the realism of the NPZ simulation. Tuna recruitment and

877

biomass simulated by SEPODYM qualitatively agreed with the estimates from the MULTIFAN-

878

CL model, and predicted catches by grid cell and month were well correlated with observed

879

catches. Most recruitment was predicted to occur in the western and central Pacific region, with

880

large variability caused by El Nino versus La Nina years. The out-of-phase dynamics of

881

simulated primary production between the western and central Pacific predicted by the NPZ

882

model was also seen in the SEAPOP-simulated tuna recruitment. They emphasized the

883

importance of a carefully constructed and evaluated NPZ model. The SEAPOP model is now

884

being applied to small pelagic species in the Humboldt system (Gaspar and Lehodey, abstract 6 ).

885 886

Shin and Cury (2004) described a general simulator of fish communities (OSMOSE) that uses an

887

individual-based size-based approach, and Shin et al. (2004) applied the model to the Benguela

888

Current system. They used the super-individual approach, with each super-individual assumed to

889

represent a school of identical fish. A distinction was made in the model between non-piscivorous

890

and piscivorous behaviour, with species assigned a behaviour based on stage or age. In the

891

Benguela application, 12 fish species were simulated on a 40 cell by 40 cell horizontal grid using

892

a 6-month time step. OSMOSE does not explicitly use the output of hydrodynamic or NPZ

893

models, but rather represents the prey field effects by specifying a system-wide carrying capacity

894

for non-piscivorous fish biomass. When total biomass of non-piscivorous species biomass

895

exceeded the carrying capacity, then mortality was imposed on all non-piscivorous individuals on

896

the grid, disproportionately on age-0 versus older individuals, until the total biomass fell below

897

the carrying capacity. Piscivorous stages consumed prey species if they co-occurred together in

898

the same spatial cell and if the prey were vulnerable based on predator to prey size ratios. All fish

6

Gaspar, P. and Lehodey, P. Application of a spatial Eulerian ecosystem and population dynamic model (SEAPODYM) to small pelagic fish: Modelling approach and preliminary tests. International Conference on The Humboldt Current System: Climate, ocean dynamics, ecosystem processes, and fisheries. Lima, Peru, November 27 December 1, 2006.

31

899

species grew in length according to von Bertalanffy equations, with the growth of piscivorous

900

species predicted by the von Bertalanffy equation affected by a predation efficiency computed for

901

each super-individual from its present consumption rate. Piscivorous super-individuals moved to

902

their neighbouring cell that had the highest prey biomass that was vulnerable to them. In addition

903

to predation, there were terms for starvation and harvesting mortality. Reproduction closed the

904

life cycle by initiating new super-individuals based on total egg production computed annually

905

from the mature female spawners; larval and juvenile survival determined subsequent

906

recruitment.

907 908

Shin et al. (2004) performed a series of 200-year simulations with increased fishing mortality

909

rates on selected species and compared predicted responses to those from a comparably

910

constructed EcoSim model. Predicted biomass of each species in OSMOSE was averaged over

911

the last 100 years of each simulation, and compared to the average biomass under the reference or

912

baseline simulation. Increased fishing morality on sardine (Sardinops sagax), anchovy (Engraulis

913

encrasicolus), and round herring (Etrumeus whiteheadi) showed that sardine would be the first to

914

collapse, anchovy would collapse second, and round herring were highly resistant because of the

915

small initial value of their fishing mortality rate. The biomass of some species, such as chub

916

mackerel (Scomber japonicus), increased due to relaxed competition. Increased fishing on Cape

917

hake (Merluccius capensis) also showed the expected decrease in hake biomass and a general

918

increase in the biomasses of hake’s competitors. Both sets of increased fishing mortality

919

simulations were compared to similar simulations preformed with an EcoSim version of the

920

Benguela Current system and, at the qualitative level (increase or decrease), both models

921

generated similar responses. While OSMOSE does not explicitly use the output of a NPZ model,

922

with some creativity, one could perhaps assume how climate change would affect prey and one

923

could adjust the carrying capacity appropriately. An ongoing effort is attempting to link an NPZ

924

model to OSMOSE so that the growth of non-piscivorous individuals can be modelled

925

dynamically (Y. Shin, personal communication).

926 927

The NEMURO (North Pacific Ecosystem Model for Understanding Regional Oceanography)

928

family of models was a milestone in an ongoing large international collaboration focused on the

929

development of standard NPZ model for application to the North Pacific, and the coupling of fish

32

930

growth and population dynamics models to this standard NPZ model. The biological food web

931

represented in NEMURO was fairly detailed with two groups of phytoplankton and three groups

932

of zooplankton, plus the usual nitrogen, silicate, and detrital recycling dynamics. Using this

933

common formulation for the NPZ, several models and applications were developed (Werner et

934

al., 2007). Of particular interest here are the spin-offs in which NEMURO was imbedded in a 3-

935

dimensional hydrodynamic model configured for the North Pacific, a version that coupled an age-

936

structured fish model to the NPZ (termed NEMURO.FISH), and the latest incarnation

937

(NEMURO.SAN) which is an individual-based, full life cycle, spatially-explicit model of sardine

938

and anchovy interactions.

939 940

NEMURO.FISH dynamically couples an adult fish bioenergetics-based population dynamics

941

model to the NEMURO NPZ model. The coupled models have been configured for Pacific

942

herring (Clupea harengus pallasii) on the west coast of Vancouver Island (Megrey et al., 2007)

943

and for Pacific saury (Cololabis saira) off Japan (Ito et al., 2007). The herring application uses an

944

age-structured approach, with bioenergetics used to describe the changes in the average body

945

weight of individuals in each age class and mortality rates used to describe the changes in the

946

numbers in each age class. Recruitment was knife-edge and computed from spawning biomass

947

and environmental variables (SST, air temperature, and North Pacific Pressure Index) using a

948

spawner-recruit relationship. New recruits become the new youngest age class. The dynamics of

949

the three zooplankton groups in the NEMURO model determine the consumption rate in the fish

950

bioenergetics model through a multispecies functional response formulation. Herring

951

consumption affects the zooplankton through predation mortality, and fish egestion and excretion

952

contribute to the nitrogen dynamics.

953 954

Using the NEMURO.FISH model of herring, Megrey et al. (2007) presented baseline simulations

955

of herring weight-at-age and population dynamics for Vancouver Island, Canada, and Rose et al.

956

(in press) used the model to examine how climate change could affect herring growth and

957

population dynamics. Rose et al. performed simulations that mimicked the conditions in each of

958

the four documented climate regimes in the North Pacific (1962-1976; 1977-1988; 1989-1997;

959

1998-2002). Climate regimes differed in the values assumed for environmental variables used in

960

the spawner-recruit relationship, and in the water temperature, mixed layer depth, and nutrient

33

961

influx rate used by the NPZ model. In agreement with general opinion and with the herring data

962

from West Coast Vancouver Island, model predicted estimates of weight-at-age, recruitment, and

963

spawning stock biomass were highest in regime 1 (1962-1976), intermediate in regime 2 (1977-

964

1988), and lowest in regime 3 (1989-1999). The regime effect on weights-at-age was a mix of

965

recruitment effects and lower trophic level effects that varied in direction and magnitude among

966

the four regimes.

967 968

Isolating the growth component of NEMURO.FISH, Rose et al. (2007) used the output of the 3-

969

dimensional NEMURO for the North Pacific and simulated weight-at-age (not population

970

dynamics) for the west coast Vancouver Island, Prince William Sound, and Bering Sea regions

971

for 1948-2002. The NEMURO application was a 3-D implementation for the Northern Pacific

972

(Aita et al., 2007). The NEMURO-3D simulation represented the NPZ dynamics for 1948 to

973

2002 using, as much possible, observed data for driving variables. The output of the NEMURO-

974

3D simulation at the three locations, averaged over cells in the top 50 m, was used as input to the

975

bioenergetics model and daily growth of herring was simulated for 1948 to 2000. Rose et al.

976

applied the sequential t-test analysis to detect regime shifts (STARS) algorithm (Rodionov and

977

Overland, 2005) to the simulated temperatures, zooplankton, and herring growth rate (annual

978

change in weight between ages 3 and 4) to identify statistical shifts in their average values. All

979

three locations showed shifts in simulated herring growth rate around the 1977 regime shift.

980

While the NEMURO-3D output showed warming temperatures at all three locations beginning in

981

the late 1970’s, herring growth was predicted to decrease in west coast Vancouver Island and

982

Prince William Sound, and to increase in the Bering Sea. Interannual variation in zooplankton

983

densities caused the time series response in herring growth for west coast Vancouver Island.

984

Temperature and zooplankton densities both played roles in Prince William Sound and the Bering

985

Sea herring growth responses, with zooplankton dominating the response for Prince William

986

Sound and temperature dominating the response for Bering Sea.

987 988

NEMURO.SAN is under development and extends the NEMURO approach to simulating sardine

989

and anchovy as individuals on a 2-dimensional spatial grid of cells (Rose et al., abstract 7 ). The

7

Rose, K. A., Agostini, V. N., Jacobson, L., van der Lingen, C., Lluch-Cota, S. E., Ito, S., Megrey, B. A., Kishi, M. J., Takasuka, A., Barange, M., Werner, F. E., Shin, Y., Cubillos, L., Yamanaka, Y. and Wei, H. Towards coupling

34

990

vertical dimension is represented as the volume of water in each cell based on the volume of the

991

water above the mixed layer depth. Each year, recruits are computed from spawning biomass at a

992

point in time and the recruits, after a suitable time delay, are slowly introduced over the next year

993

as new model individuals on the grid. Growth, mortality, and movement of individuals are

994

evaluated daily. Positions of individuals are tracked in continuous x and y space, and each day

995

their cell is determined and they experience the conditions in that cell. Alternative approaches to

996

movement, including the use of Railsback and Humston approaches, are being investigated. How

997

the spatial (among cells) and temporal (daily or monthly) variation in the mixed layer depth,

998

nutrient influx, and other inputs to NPZ portion are specified allows for flexibility in configuring

999

NEMURO.SAN to different locations. To date, 100-year simulations have been performed in an

1000

exploratory mode for a version configured to roughly resemble the California Current system.

1001

Alternative hypotheses about climate conditions can be specified via changing the inputs to

1002

NEMURO, and then using the coupled models to predict the long-term responses of the anchovy

1003

and sardines in terms of their growth, survival, and spatial distributions.

1004 1005

4 Biophysical models and climate change

1006 1007

In this section we discuss the use of biophysical models of small pelagic fish in the context of

1008

predicting the effects of climate change. Harley et al. (2006) recently reviewed the potential

1009

impacts of climate change on coastal marine ecosystems. They discussed ecological responses to

1010

climate change at the individual, population, and community levels. We envision most analyses

1011

using biophysical models being single species analyses. The early life stage models focus on a

1012

single species, and while some of the full life cycle models include multiple species, there is still

1013

great uncertainty in how to represent the food web dynamics of the upper trophic level species

1014

(Rose and Sable, 2007). Despite the recognition of the importance of food web interactions and

1015

the push towards ecosystem-based fisheries management (Rose and Sable, 2007), community-

1016

level responses to climate change that require multi-species simulations will likely remain in the

1017

demonstration mode, rather than for management decision-making, for the foreseeable future. We

sardine and anchovy to the NEMURO lower trophic level model. North Pacific Marine Science Organization (PICES) 15th Annual Meeting, Yokohama, Japan, October 13-22, 2006.

35

1018

envision most analyses in the near term will focus mainly on the early life stages of key species,

1019

with some analysis using single-species full life cycle models for well-studied locations and

1020

theoretical-oriented analyses using the multispecies full life cycle models.

1021 1022

Among the factors reviewed by Harley et al. (2006), the ecological responses to changes in

1023

circulation, temperature, and productivity are relatively easy to investigate using biophysical

1024

models. Indeed, changes in circulation would induce changes in advective and dispersive

1025

transport, a process used directly in the early life stage models, and used directly or indirectly in

1026

the adult-based models. Some of the reviewed models used the direct effects of transport to at

1027

least influence the movement of adult fish, and many of the reviewed models used the effects of

1028

transport indirectly via transport’s effects on temperature, salinity, and prey fields that affected

1029

the growth, mortality, reproduction, and movement rates of adult fish. Harley et al. also discussed

1030

the potential shifts (vertical and horizontal) in species distributions that could occur under climate

1031

change. Initial conditions in early life stage models typically use data based on observed egg

1032

distributions. Shifts in egg distributions induced by climate change could be hypothesized and

1033

simulations performed to explore potential consequences on early life stage transport, growth,

1034

and survival.

1035 1036

To our knowledge, the only example of biophysical model used in the context of climate change

1037

comes from the recent study of Vikebø et al. (2007). In the Barents Sea off Norway, they used the

1038

ROMS hydrodynamic model forced by a global climate model in which the river runoff was

1039

increased by a factor three over the current value, causing the thermohaline circulation to slow

1040

down. Then they used a biophysical model of the early life stages of Arcto-Norwegian cod

1041

(Gadus morhua) to study the impact of the anomalous circulation and ocean temperature on

1042

transport and growth of cod larvae and juveniles. They showed clear differences between the drift

1043

and growth patterns obtained under the current and the reduced thermohaline circulation (Figure

1044

3).

1045 1046

Figure 3

1047

36

1048

We have more confidence in the early life stage models for simulating climate change scenarios

1049

because they are more tightly coupled to the physics and NPZ, and because early life stages of

1050

fish tend to be the focus of field and modelling studies in the marine ecosystems. Early life stages

1051

can be measured and they are often the focus as part of the search for recruitment indices

1052

(Kendall and Duker, 1998). Imposing climate change scenarios would seem to be possible almost

1053

from first principles because the early life stage models are tightly coupled to the hydrodynamic

1054

and NPZ models. Simulating climate change scenarios becomes more complicated for the adult

1055

models because many of the adult models used surrogates for the physics and prey outputs of the

1056

hydrodynamic and NPZ models as their inputs (e.g., carrying capacity). This implies another step

1057

is required to convert the hydrodynamics and NPZ outputs into the variables needed by the adult

1058

models.

1059 1060

Predicting responses of early life stages is necessary but not sufficient to address many but not all

1061

of the climate change issues. Many issues require that the predictions of early life stage models be

1062

related to population dynamics and health, and some questions require the simulation of the

1063

responses of the adult life stages. One challenge is to balance the use of early life stage, adult

1064

only, and full life cycle modelling approaches to ensure predicted responses are most relevant to

1065

the scientific questions and to management issues. There are also practical computing challenges.

1066

For example, the numerical considerations that arose with using super-individuals with early life

1067

stage models become more complicated with full life cycle models, as these models often include

1068

density-dependent effects and require new sets of individuals be introduced each year of their

1069

multi-year simulations. Also, full life cycle models typically simulate decades in order for the full

1070

effects of environmental and climate changes to be manifested in the population response. Yet,

1071

hydrodynamic and NPZ models typically simulate a few years. How one creates realistic long-

1072

term scenarios for use as input to the full life cycle model from the relatively limited number of

1073

hydrodynamic and NPZ results can affect predicted responses to climate change.

1074 1075

Perhaps the greatest challenge relates to the assumption that the predicted responses of the

1076

hydrodynamic and NPZ models under climate change scenarios are sufficiently accurate and

1077

precise, and on the correct spatial and temporal scales to be used as inputs to the fish models.

1078

There is still disagreement about how some important driving variables to the hydrodynamic and

37

1079

NPZ model (e.g., boundary conditions, precipitation, wind) will change under a given climate

1080

change scenario. Furthermore, climate change scenarios, even after agreed upon, can push the

1081

hydrodynamic and NPZ models beyond their calibrated and validated domain. Also, most climate

1082

change simulation experiments have been conducted at a large spatial scale (e.g., the entire

1083

Southern Ocean, Russell et al., 2006), whereas the early life stage and adult models tend to

1084

operate at smaller scales. Regional-level climate change simulation experiments should become

1085

increasingly available in the near future (e.g., for the California Current system, Auad et al.,

1086

2006). Meanwhile, creative use of long-term time series of historical hydrodynamic interannual

1087

simulations should be melded with climate change scenarios to ensure realistic outputs of the

1088

hydrodynamic and NPZ models that act as inputs to the fish models.

1089 1090

Models such as the Mullon et al. (2002) model could be used to derive expected egg distributions

1091

under climate change. For example, simulation 4 of Mullon et al. (2002) was re-run with

1092

hydrodynamic model outputs using weekly wind forcing from ERS satellites (run C in Blanke et

1093

al., 2002) to study how simulated selected spawning patterns would change according to

1094

evolutionary constraints that involve a lethal temperature threshold for the transported particles.

1095

Two main spatio-temporal spawning patterns emerged depending on the value of the temperature

1096

threshold. A cooler threshold selected spawning in the Central Agulhas Bank in July–September

1097

(Figure 4a), while a warmer threshold resulted in the selection of spawning in the Eastern

1098

Agulhas Bank in October–January (Figure 4c); an intermediate temperature threshold resulted in

1099

a mix of both patterns (Figure 4b). Assuming that the spawning pattern of Figure 4b corresponds

1100

to the current situation, and that climatic change would only result in homogeneous increase of

1101

sea surface temperature, one could infer that spawning would shift more towards the pattern

1102

shown in Figure 2a (i.e., cooler threshold mimics warmer water). These results can then be used

1103

as inputs to early life stage biophysical models.

1104 1105

Figure 4

1106 1107

Simulating geographic shifts in adult life stages will be more problematic with biophysical adult

1108

models. The use of adult models for simulating geographic shifts is limited because the distances

1109

that could be moved by adult fish in response to climate change would likely exceed the spatial

38

1110

domain of most of the models. A few of the adult-based fish models used very large spatial

1111

domains (e.g., Lehodey et al., 2003), but whether the biological aspects of these models are

1112

sufficiently general to allow for prediction of geographic shifts needs to be tested. This was

1113

partially addressed in the SEPODYM model because they had to deal with the highly migratory

1114

tuna. Whether the current models of small pelagic fish are sufficiently robust to simulate large-

1115

scale shifts in distribution remains to be determined. To date, more statistically-oriented empirical

1116

approaches have been used to predict geographic shifts in the distributions of adult fish and other

1117

taxa in response to climate change (e.g., Rahel, 2002; Schmitz et al., 2003). Allowing for the

1118

capability for predicting geographic shifts should be considered as present biophysical models are

1119

expanded and new models are developed.

1120 1121

Biophysical models of fish early life stages could likely be improved by using spawning and

1122

nursery habitats defined by environmental variables rather than simply being geo-referenced (P.

1123

Fréon, personal communication). Characterizations of spawning habitats based upon

1124

environmental variables such as temperature and salinity have been performed for anchovy and

1125

sardine in California (Lynn, 2003), the southern Benguela (Twatwa et al., 2005), the Bay of

1126

Biscay (Planque et al., 2007), and elsewhere (Castro et al., 2005; van der Lingen et al., 2005).

1127

Relating spawning habitat to environmental conditions would allow for the simulation of an

1128

“environmental homing” reproduction strategy as opposed to a “natal homing” strategy (Cury,

1129

1994), and allow investigation of climate-driven changes in the habitat and their consequences on

1130

early life stages dynamics. Some caution is appropriate for such a “climate envelope” approach

1131

(Davis et al., 1998) because the characterizations by environmental variables may themselves

1132

change in response to climate change due to adaptation, acclimation, and changes in the food

1133

web.

1134 1135

Early life stage models are ready for species-specific and site-specific analysis of climate change

1136

effects of small pelagic fish in upwelling systems. Some of the details of such applications need

1137

careful attention, especially the matching of the spatial and temporal scales of the hydrodynamic

1138

and NPZ models with the fish models, and how growth, mortality, behaviour-based movement

1139

and trophic feedbacks are represented (Runge et al., 2005). We see an accelerating trend from

1140

models of “hydrodynamics and simple behaviors” to models of “hydrodynamics and dynamic

39

1141

prey”. It may be possible to include invertebrate predators in the models, but it will remain

1142

difficult to include fish predators of eggs and larvae. A large uncertainty may be getting

1143

agreement about how climate change will affect hydrodynamic outputs that are then used as

1144

inputs to the fish models, and obtaining these outputs on the spatial and temporal scales needed

1145

by the biophysical fish models.

1146 1147

Biophysical models that include the adult life stages of fish is an area of high research interest but

1148

will likely be used for general analysis or application to a few, extremely well-studied species and

1149

locations. The uncertainty in how to model the behavioural aspects of movement will limit the

1150

development of a generally agreed upon modelling approach (i.e., physics-based) that has helped

1151

advance the early life stage models. New measurement methods are becoming available that

1152

should provide the empirical basis for evaluating the alternative movement options (Cooke et al.,

1153

2004). Sugden and Pennisi (2006) recently introduced a special section in Science on “movement

1154

ecology”. Continued efforts on adult models is necessary in order to get to “end-to-end” (physics

1155

to fish), full life cycle models capable of addressing the long-term consequences of climate

1156

change on fish and fisheries.

1157

40

1157

Figures

1158

1159 1160

Figure 1: A schematic view of the approach generally used for implementing biophysical models

1161

of marine species early life history (modified from Hermann et al., 2001). An hydrodynamic

1162

1164

model (or an hydrodynamic-biogeochemical coupled model) provides three-dimensional dynamic r fields of current velocities u , temperature T , and other variables (e.g., phytoplankton biomass r P ), to an individual-based model that tracks location x , length L and other variables of interest

1165

for a collection of individuals i over time t .

1163

1166

41

1166

(a)

(b)

(c)

(d)

1167

Figure 2: Crab larvae concentrations (a) observed and (b–d) obtained from a biophysical model

1168

that included (b) diel vertical migration for larvae and river outflow (c) only diel vertical

1169

migration for larvae (d) only river outflow. Reproduced from Peliz et al. (in press).

1170

42

1170

(a)

(b)

1171

Figure 3: Simulated distribution and wet weight (colour scale, in mg) of 2-4 month old juvenile

1172

cod obtained from a biophysical model using hydrodynamics under (a) current conditions and (b)

1173

under reduced thermohaline circulation. Reproduced from Vikebø et al. (2007).

1174

43

1174

(a)

(b)

(c)

1175

Figure 4: Spatio-temporal spawning pattern after 200 generations obtained from a biophysical

1176

model that used 100 000 particles and a lethal temperature threshold of (a) 14°C (like simulation

1177

4 in Mullon et al., 2002) (b) 15°C (c) 16°C.

1178

44

1178

References

1179 1180

Acreman, M. and Dunbar, M. J. (2004). Defining environmental river flow requirements – a

1181

review. Hydrology and Earth System Sciences, 8, 861-876.

1182 1183

Adamack, A. (2007). Predicting water quality effects on bay anchovy (Anchoa mitchilli) growth

1184

and production in Chesapeake Bay: linking water quality and individual-based fish models. Ph.D.

1185

thesis, Louisiana State University, Baton Rouge, LA, USA.

1186 1187

Ådlandsvik, B., Gundersen, A. C., Nedreaas, K. H., Stene, A. and Albert, O. T. (2004).

1188

Modelling the advection and diffusion of eggs and larvae of Greenland halibut (Reinhardtius

1189

hippoglossoides) in the north-east Arctic. Fisheries Oceanography, 13, 403-415.

1190 1191

Aita, M. N., Yamanaka, Y. and Kishi, M. J. (2007). Interdecadal variation of the lower trophic

1192

ecosystem in the northern Pacific between 1948 and 2002, in a 3-D implementation of the

1193

NEMURO model. Ecological Modelling, 202, 81-94.

1194 1195

Allain, G., Petitgas, P., Grellier, P. and Lazure, P. (2003). The selection process from larval to

1196

juvenile stages of anchovy (Engraulis encrasicolus) in the Bay of Biscay investigated by

1197

Lagrangian simulations and comparative otolith growth. Fisheries Oceanography, 12, 407-418.

1198 1199

Allain, G. (2004). Modélisation biophysique pour la prévision du recrutement: couplage

1200

stochastique d'un modèle individu-centré de croissance larvaire avec un modèle hydrodynamique

1201

3D pour développer un indice de recrutement de l'anchois dans le golfe de Gascogne. Ph.D.

1202

thesis, Ecole Nationale Supérieure Agronomique de Rennes, France.

1203 1204

Allain, G., Petitgas, P. and Lazure, P. (in press). Biophysical modelling of larval drift, growth and

1205

survival for the prediction of anchovy recruitment in the Bay of Biscay (NE Atlantic). Fisheries

1206

Oceanography.

1207

45

1208

Anderson, J. J. (2002). An agent-based event driven foraging model. Natural Resource Modeling,

1209

15, 55-82.

1210 1211

Andrews, J. M., Gurney, W. S. C., Heath, M. R., Gallego, A., O'Brien, C. M., Darby, C. and

1212

Tyldesley, G. (2006). Modelling the spatial demography of Atlantic cod (Gadus morhua) on the

1213

European continental shelf. Canadian Journal of Fisheries and Aquatic Sciences, 63, 1027-1048.

1214 1215

Auad, G., Miller, A. and Di Lorenzo, E. (2006). Long-term forecast of oceanic conditions off

1216

California and their biological implications. Journal of Geophysical Research-Oceans, 111,

1217

C09008, doi:09010.01029/02005JC003219.

1218 1219

Ault, J. S., Luo, J., Smith, S. G., Serafy, J. E.,Wang, J. D., Humston, R. and Diaz, G. A. (1999).

1220

A spatial dynamic multistock production model. Canadian Journal of Fisheries and Aquatic

1221

Sciences 56 (Suppl.1):4-25.

1222 1223

Bakun, A. (1996). Patterns in the ocean. Ocean processes and marine population dynamics.

1224

University of California Sea Grant, California, USA, in cooperation with Centro de

1225

Investigaciones Biologicas de Noroeste, La Paz, Baja California Sur, Mexico.

1226 1227

Bartsch, J., Brander, K., Heath, M., Munk, P., Richardson, K. and Svendsen, E. (1989).

1228

Modelling the advection of herring larvae in the North Sea. Nature, 340, 632-636.

1229 1230

Bartsch, J. (1993). Application of a circulation and transport model system to the dispersal of

1231

herring larvae in the North Sea. Continental Shelf Research, 13, 1335-1361.

1232 1233

Bartsch, J. and Knust, R. (1994a). Simulating the dispersion of vertically migrating sprat larvae

1234

(Sprattus sprattus (L.)) in the German Bight with a circulation and transport model system.

1235

Fisheries Oceanography, 3, 92-105.

1236 1237

Bartsch, J. and Knust, R. (1994b). Predicting the dispersion of sprat larvae (Sprattus sprattus

1238

(L.)) in the German Bight. Fisheries Oceanography, 3, 293-296.

46

1239 1240

Bartsch, J. and Coombs, S. H. (2001). An individual-based growth and transport model of the

1241

early life-history stages of mackerel (Scomber scombrus) in the eastern North Atlantic.

1242

Ecological Modelling, 138, 127-141.

1243 1244

Bartsch, J. and Coombs, S. H. (2004). An individual-based model of the early life history of

1245

mackerel (Scomber scombrus) in the eastern North Atlantic, simulating transport, growth and

1246

mortality. Fisheries Oceanography, 13, 365-379.

1247 1248

Bartsch, J., Reid, D. and Coombs, S. H. (2004). Simulation of mackerel (Scomber scombrus)

1249

recruitment with an individual-based model and comparison with field data. Fisheries

1250

Oceanography, 13, 380-391.

1251 1252

Bartsch, J. (2005). The influence of spatio-temporal egg production variability on the modelled

1253

survival of the early life history stages of mackerel (Scomber scombrus) in the eastern North

1254

Atlantic. Ices Journal of Marine Science, 62, 1049-1060.

1255 1256

Baumann, H., Hinrichsen, H.-H., Möllmann, C., Köster, F. W., Malzahn, A. M. and Temming, A.

1257

(2006). Recruitment variability in Baltic Sea sprat (Sprattus sprattus) is tightly couplesd to

1258

temperature and transport patterns affecting the larval and early juvenile stages. Canadian

1259

Journal of Fisheries and Aquatic Sciences, 63, 2191-2201.

1260 1261

Blackwell, P. G. (1997) Random diffusion models for animal movement. Ecological Modelling,

1262

100, 87-102.

1263 1264

Blanke, B., Roy, C., Penven, P., Speich, S., McWilliams, J. and Nelson, G. (2002). Linking wind

1265

and interannual upwelling variability in a regional model of the southern Benguela. Geophysical

1266

Research Letters, 29, 2188, doi:2110.1029/2002GL015718.

1267 1268

Botsford, L. W., Moloney, C. L., Hastings, A., Largier, J. L., Powell, T. M., Higgins, K. and

1269

Quinn, J. F. (1994). The influence of spatially and temporally varying oceanographic conditions

47

1270

on meroplanktonic metapopulations. Deep Sea Research Part II: Topical Studies in

1271

Oceanography, 41, 107-145.

1272 1273

Brickman, D., Shackell, N. L. and Frank, K. T. (2001). Modelling the retention and survival of

1274

Browns Bank haddock larvae using an early life stage model. Fisheries Oceanography, 10, 284-

1275

296.

1276 1277

Brown, C. A., Jackson, G. A., Holt, S. A. and Holt, G. J. (2005). Spatial and temporal patterns in

1278

modeled particle transport to estuarine habitat with comparisons to larval fish settlement patterns.

1279

Estuarine, Coastal and Shelf Science, 64, 33-46.

1280 1281

Carr, S. D., Capet, X. J., McWilliams, J. C., Pennington, J. T. and Chavez, F. P. (in press). The

1282

influence of diel vertical migration on zooplankton transport and recruitment in an upwelling

1283

region: estimates from a coupled behavioral-physical model. Fisheries Oceanography.

1284 1285

Castro, L. R., Fréon, P., van der Lingen, C. D. and Uriarte, A. (2005). Report of the SPACC

1286

meeting on small pelagic fish spawning habitat dynamics and the daily egg production method

1287

(DEPM). GLOBEC Report 22, xiv.

1288 1289

Clark, R. A., Fox, C. J., Viner, D. and Livermore, M. (2003). North Sea cod and climate change -

1290

modelling the effects of temperature on population dynamics. Global Change Biology, 9, 1669-

1291

1680.

1292 1293

Cooke, S. J., Hinch, S. G., Wikelski, M., Andrews R. D., Kuchel, L. J., Wolcott, T. G. and Butler,

1294

P. J. (2004). Biotelemetry: a mechanistic approach to ecology. Trends in Ecology and Evolution,

1295

19, 334-343.

1296 1297

Cury, P. (1994). Obstinate nature: an ecology of individuals. Thoughts on reproductive behavior

1298

and biodiversity. Canadian Journal of Fisheries and Aquatic Sciences, 51, 1664-1673.

1299

48

1300

Cury, P., Bakun, A., Crawford, R. J. M., Jarre, A., Quiñones, R. A., Shannon, L. J. and Verheye,

1301

H. M. (2000). Small pelagics in upwelling systems: patterns of interaction and structural changes

1302

in "wasp-waist" ecosystems. Ices Journal of Marine Science, 57, 603-618.

1303 1304

Davis, A. J., Jenkinson, L. S., Lawton, J. H., Shorrocks, B. and Wood, S. (1998). Making

1305

mistakes when predicting shifts in species range in response to global warming. Nature, 391,

1306

783-786.

1307 1308

Demarcq, H., Barlow, R. G. and Hutchings, L. (in press). Application of a chlorophyll index

1309

derived from satellite data to investigate the variability of phytoplankton in the Benguela

1310

ecosystem. African Journal of Marine Science.

1311 1312

Dopolo, M. T., van der Lingen, C. D. and Moloney, C. L. (2005). Stage-dependent vertical

1313

distribution of pelagic fish eggs on the western Agulhas Bank, South Africa. African Journal of

1314

Marine Science, 27, 249-256.

1315 1316

Edwards, K. P., Hare, J. A., Werner, F. E. and Blanton, B. O. (2006). Lagrangian circulation on

1317

the southeast US continental shelf: implications for larval dispersal and retention. Continental

1318

Shelf Research, 26, 1375-1394.

1319 1320

Fach, B. A. and Klinck, J. M. (2006). Transport of Antarctic krill (Euphausia superba) across the

1321

Scotia Sea. Part I: Circulation and particle tracking simulations. Deep-Sea Research Part I-

1322

Oceanographic Research Papers, 53, 987-1010.

1323 1324

Field, J. C., Francis, R. C. and Aydin, K. (2006). Top-down modeling and bottom-up dynamics:

1325

linking a fisheries-based ecosystem model with climate hypotheses in the northern California

1326

Current. Progress in Oceanography, 68, 238-270.

1327 1328

Fiksen, O. and Slotte, A. (2002). Stock-environment recruitment models for Norwegian spring

1329

spawning herring (Clupea harengus). Canadian Journal of Fisheries and Aquatic Sciences, 59,

1330

211-217.

49

1331 1332

Fiksen, O., Jørgensen, C., Kristiansen, T., Vikebø, F. and Huse, G. (in press). Linking behavioral

1333

ecology and oceanography: larval behaviour determines growth, mortality and dispersal. Marine

1334

Ecology-Progress Series.

1335 1336

Fournier, D. A., Hampton, J. and Sibert, J. R. (1998). MULTIFAN-CL : a length-based, aged-

1337

structured model for fisheries stock assessment, with application to South Pacific albacore,

1338

Thunnus Alalunga. Canadian Journal of Fisheries and Aquatic Sciences, 55, 2105- 2116.

1339 1340

Fox, C. J., McCloghrie, P., Young, E. F. and Nash, R. D. M. (2006). The importance of

1341

individual behaviour for successful settlement of juvenile plaice (Pleuronectes platessa L.): a

1342

modelling and field study in the eastern Irish Sea. Fisheries Oceanography, 15, 301-313.

1343 1344

Fulton, E. A., Smith, A. D. M. and Johnson, C. R. (2004a). Biogeochemical marine ecosystem

1345

models I: IGBEM - a model of marine bay ecosystems. Ecological Modelling, 174, 267-307.

1346 1347

Fulton, E. A., Parslow, J. S., Smith, A. D. M. and Johnson, C. R. (2004b). Biogeochemical

1348

marine ecosystem models II: the effect of physiological detail on model performance. Ecological

1349

Modelling, 173, 371-406.

1350 1351

Goodwin, R. A., Nestler, J. M., Anderson, J. J., Weber, L. J. and Loucks, D. P. (2006).

1352

Forecasting 3-D fish movement behavior using a Eulerian-Lagrangian-agent method (ELAM).

1353

Ecological Modelling, 192, 197-223.

1354 1355

Grimm, V., Revilla, E., Berger, U., Jeltsch, F., Mooij, W. M., Railsback, S. F., Thulke, H. H.,

1356

Weiner, J., Wiegand, T. and DeAngelis, D. L. (2005). Pattern-oriented modeling of agent-based

1357

complex systems: Lessons from ecology. Science, 310, 987-991.

1358 1359

Guay, J. C., Boisclair, D., Leclerc, M., Lapointe, M. and Legendre, P. (2000). Development and

1360

validation of numerical habitat models for juveniles of Atlantic salmon (Slamo slari). Canadian

1361

Journal of Fisheries and Aquatic Sciences, 57, 2065-2075.

50

1362 1363

Guizien, K., Brochier, T., Duchene, J. C., Koh, B. S. and Marsaleix, P. (2006). Dispersal of

1364

Owenia fusiformis larvae by wind-driven currents: turbulence, swimming behaviour and

1365

mortality in a three-dimensional stochastic model. Marine Ecology-Progress Series, 311, 47-66.

1366 1367

Gutiérrez, O. Q., Marinone, S. G. and Parés-Sierra, A. (2004). Lagrangian surface circulation in

1368

the Gulf of California from a 3D numerical model. Deep-Sea Research Part II-Topical Studies in

1369

Oceanography, 51, 659-672.

1370 1371

Hanson, P. C., Johnson, T. B., Schindler, D. E. and Kitchell, J. F. (1997). Fish bioenergetics 3.0.

1372

University of Wisconsin Sea Grant Institute, WISCU-T-97-001, Madison, Wisconsin.

1373 1374

Hao, W., Jian, S., Ruijing, W., Lei, W. and Yi'An, L. (2003). Tidal front and the convergence of

1375

anchovy (Engraulis japonicus) eggs in the Yellow Sea. Fisheries Oceanography, 12, 434-442.

1376 1377

Harley, C. D. G., Hughes, A. R., Hultgren, K. M., Miner, B. G., Sorte, C. J. B., Thornber, C. S.,

1378

Rodriguez, L. F., Tomanek, L. and Williams, S. L. (2006). The impacts of climate change in

1379

coastal marine systems. Ecology Letters, 9, 228-241.

1380 1381

Haza, A. C., Piterbarg, L. I., Martin, P., Özgökmen, T. M. and Griffa, A. (in press). A Lagrangian

1382

subgridscale model for particle transport improvement and application in the Adriatic Sea using

1383

the Navy Coastal Ocean Model. Ocean Modelling, doi:10.1016/j.ocemod.2006.1010.1004.

1384 1385

Heath, M., Scott, B. and Bryant, A. D. (1997). Modelling the growth of herring from four

1386

different stocks in the North Sea. Journal of Sea Research, 38, 413-436.

1387 1388

Heath, M. and Gallego, A. (1997). From the biology of the individual to the dynamics of the

1389

population: bridging the gap in fish early life studies. Journal of Fish Biology, 51, 1-29.

1390

51

1391

Heath, M., Zenitani, H., Watanabe, Y., Kimura, R. and Ishida, M. (1998). Modelling the dispersal

1392

of larval Japanese sardine, Sardinops melanostictus, by the Kuroshio Current in 1993 and 1994.

1393

Fisheries Oceanography, 7, 335-346.

1394 1395

Hermann, A. J., Hinckley, S., Megrey, B. A. and Napp, J. M. (2001). Applied and theoretical

1396

considerations for constructing spatially explicit individual-based models of marine larval fish

1397

that include multiple trophic levels. Ices Journal of Marine Science, 58, 1030-1041.

1398 1399

Hinckley, S. (1999). Biophysical mechanisms underlying the recruitment process in walleye

1400

pollock (Theragra chalcogramma). Ph.D. thesis, University of Washington, USA.

1401 1402

Hinrichsen, H. H., Möllmann, C., Voss, R., Köster, F. W. and Kornilovs, G. (2002). Biophysical

1403

modeling of larval Baltic cod (Gadus morhua) growth and survival. Canadian Journal of

1404

Fisheries and Aquatic Sciences, 59, 1858-1873.

1405 1406

Hinrichsen, H.-H., Kraus, G., Voss, R., Stepputis, D. and Baumann, H. (2005). The general

1407

distribution and mixing probability of Baltic sprat juvenile populations. Journal of Marine

1408

Systems, 58, 52-66.

1409 1410

Huggett, J., Fréon, P., Mullon, C. and Penven, P. (2003). Modelling the transport success of

1411

anchovy Engraulis encrasicolus eggs and larvae in the southern Benguela: the effect of spatio-

1412

temporal spawning patterns. Marine Ecology-Progress Series, 250, 247-262.

1413 1414

Humston, R., Olson, D. B. and Ault, J. S. (2004). Behavioral assumptions in models of fish

1415

movement and their influence on population dynamics. Transactions of the American Fisheries

1416

Society, 133, 1304-1328.

1417 1418

Hunter, J. R. (1977). Behavior and survival of northern anchovy Engraulis Mordax larvae.

1419

California Cooperative Oceanic Fisheries Investigations Reports, 19, 138-146.

1420

52

1421

Hunter, J. R. and Coyne, K. M. (1982). The onset of schooling in northern anchovy larvae,

1422

Engraulis mordax. California Cooperative Oceanic Fisheries Investigations Reports, 23, 246-

1423

251.

1424 1425

Huret, M., Runge, J. A., Chen, C., Cowles, G., Xu, Q. and Pringle, J. M. (in press). Dispersal

1426

modeling of fish early life stages: sensitivity with application to Atlantic cod in the western Gulf

1427

of Maine. Marine Ecology-Progress Series.

1428 1429

Huse, G. and Giske, J. (1998). Ecology in the Mare Pentium: an individual-based spatio-temporal

1430

model for fish with adapted behaviour. Fisheries Research, 37, 163-178.

1431 1432

Ito, S., Megrey, B. A., Kishi, M. J., Mukai, D., Kurita, Y., Ueno, Y. and Yamanaka, Y. (2007).

1433

On the interannual variability of the growth of Pacific saury (Cololabis saira): a simple 3-box

1434

model using NEMURO.FISH. Ecological Modelling, 202, 174-183.

1435 1436

Jacobson, L. D., Bograd, S. J., Parrish, R. H., Mendelssohn, R. and Schwing, F. B. (2005). An

1437

ecosystem-based hypothesis for climatic effects on surplus production in California sardine

1438

(Sardinops sagax) and environmentally dependent surplus production models. Canadian Journal

1439

of Fisheries and Aquatic Sciences, 62, 1782-1796.

1440 1441

Karim, M. R., Sekine, M., Higuchi, T., Imai, T. and Ukita, M. (2003). Simulation of fish behavior

1442

and mortality in hypoxic water in an enclosed bay. Ecological Modelling, 159, 27-42.

1443 1444

Kendall, A. W. and Duker, G. J. (1998). The development of recruitment fisheries oceanography

1445

in the United States. Fisheries Oceanography, 7, 69-88.

1446 1447

Koné, V., Machu, E., Penven, P., Andersen, V., Garçon, V., Fréon, P. and Demarcq, H. (2005).

1448

Modeling the primary and secondary productions of the southern Benguela upwelling system: A

1449

comparative study through two biogeochemical models. Global Biogeochemical Cycles, 19,

1450

GB4021, doi:4010.1029/2004GB002427.

1451

53

1452

Koné, V. (2006). Modélisation de la production primaire et secondaire de l'écosystème du

1453

Benguela sud. Influence des conditions trophiques sur le recrutement des larves d'anchois. Ph.D.

1454

thesis, Université Pierre & Marie Curie (Paris VI), France.

1455 1456

Lazure, P. and Dumas, F. (in press). A 3D hydrodynamics model for applications at the regional

1457

scale (MARS-3D): application to the bay of Biscay. Advances in Water Resources.

1458 1459

Lehodey, P., Chai, F. and Hampton, J. (2003). Modelling climate-related variability of tuna

1460

populations from a coupled ocean-biogeochemical-population dynamics model. Fisheries

1461

Oceanography, 12, 483-494.

1462 1463

Lett, C., Roy, C., Levasseur, A., van der Lingen, C. D. and Mullon, C. (2006). Simulation and

1464

quantification of enrichment and retention processes in the southern Benguela upwelling

1465

ecosystem. Fisheries Oceanography, 15, 363-372.

1466 1467

Lett, C., Penven, P., Ayón, P. and Fréon, P. (2007). Enrichment, concentration and retention

1468

processes in relation to anchovy (Engraulis ringens) eggs and larvae distributions in the northern

1469

Humboldt upwelling ecosystem. Journal of Marine Systems, 64, 189-200.

1470 1471

Lett, C., Veitch, J., van der Lingen, C. D. and Hutchings, L. (in press). Assessment of an

1472

environmental barrier to transport of ichthyoplankton from the southern to the northern Benguela

1473

ecosystems. Marine Ecology-Progress Series.

1474 1475

Lough, R. G., Buckley, L. J., Werner, F. E., Quinlan, J. A. and Pehrson Edwards, K. (2005). A

1476

general biophysical model of larval cod (Gadus morhua) growth applied to populations on

1477

Georges Bank. Fisheries Oceanography, 14, 241-262.

1478 1479

Lough, R. G., Broughton, E. A., Buckley, L. J., Incze, L. S., Edwards, K. P., Converse, R.,

1480

Aretxabaleta, A. and Werner, F. E. (2006). Modeling growth of Atlantic cod larvae on the

1481

southern flank of Georges Bank in the tidal-front circulation during May 1999. Deep Sea

1482

Research Part II: Topical Studies in Oceanography, 53, 2771-2788.

54

1483 1484

Luo, J., Hartman, K. J., Brandt, S. B., Cerco, C. F. and Rippetoe, T. H. (2001). A spatially-

1485

explicit approach for estimating carrying capacity: an application for the Atlantic menhaden

1486

(Brevoortia tyrannus) in Chesapeake Bay. Estuaries, 24, 545-556.

1487 1488

Lynn, R. J. (2003). Variability in the spawning habitat of Pacific sardine (Sardinops sagax) off

1489

southern and central California. Fisheries Oceanography, 12, 541-553.

1490 1491

Marinone, S. G. (2003). A three-dimensional model of the mean and seasonal circulation of the

1492

Gulf of California. Journal of Geophysical Research-Oceans, 108, C10, 3325,

1493

doi:3310.1029/2002JC001720.

1494 1495

Marinone, S. G., Gutierrez, O. Q. and Pares-Sierra, A. (2004). Numerical simulation of larval

1496

shrimp dispersion in the Northern Region of the Gulf of California. Estuarine Coastal and Shelf

1497

Science, 60, 611-617.

1498 1499

Marta-Almeida, M., Dubert, J., Peliz, A. and Queiroga, H. (2006). Influence of vertical migration

1500

pattern on retention of crab larvae in a seasonal upwelling system. Marine Ecology-Progress

1501

Series, 307, 1-19.

1502 1503

Megrey, B. A. and Hinckley, S. (2001). Effect of turbulence on feeding of larval fishes: a

1504

sensitivity analysis using an individual-based model. Ices Journal of Marine Science, 58, 1015-

1505

1029.

1506 1507

Megrey B. A., Rose, K. A., Klumb, R. A, Hay, D. E., Werner, F. E., Eslinger, D. L. and Smith, S.

1508

L. (2007). A bioenergetics-based population dynamics model of Pacific herring (Clupea harengus

1509

pallasi) coupled to a lower trophic level nutrient-phytoplankton-zooplankton model: Description,

1510

calibration, and sensitivity analysis. Ecological Modelling, 202, 144-164.

1511

55

1512

Miller, C. B., Lynch, D. R., Carlotti, F., Gentleman, W. and Lewis, C. V. W. (1998). Coupling of

1513

an individual-based population dynamic model of Calanus finmarchicus to a circulation model

1514

for the Georges Bank region. Fisheries Oceanography, 7, 219-234.

1515 1516

Miller, D. C. M. (2006). An individual-based modelling approach to examine life history

1517

strategies of sardine (Sardinops sagax) in the southern Benguela ecosystem. Ph.D. thesis,

1518

University of Cape Town, South Africa.

1519 1520

Miller, D. C. M., Moloney, C. L., van der Lingen, C. D., Lett, C., Mullon, C. and Field, J. G.

1521

(2006). Modelling the effects of physical-biological interactions and spatial variability in

1522

spawning and nursery areas on transport and retention of sardine eggs and larvae in the southern

1523

Benguela ecosystem. Journal of Marine Systems, 61, 212-229.

1524 1525

Miller, T. J. (in press). Contribution of individual-based coupled physical biological models to

1526

understanding recruitment in marine fish populations. Marine Ecology-Progress Series.

1527 1528

Mullon, C., Cury, P. and Penven, P. (2002). Evolutionary individual-based model for the

1529

recruitment of anchovy (Engraulis capensis) in the southern Benguela. Canadian Journal of

1530

Fisheries and Aquatic Sciences, 59, 910-922.

1531 1532

Mullon, C., Fréon, P., Parada, C., van der Lingen, C. and Huggett, J. (2003). From particles to

1533

individuals: modelling the early stages of anchovy (Engraulis capensis/encrasicolus) in the

1534

southern Benguela. Fisheries Oceanography, 12, 396-406.

1535 1536

Ney, J. J. (1993). Bioenergetics modeling today: growing pains on the cutting edge. Transactions

1537

of the American Fisheries Society, 122, 736-748.

1538 1539

North, E. W., Hood, R. R., Chao, S.-Y. and Sanford, L. P. (2005). The influence of episodic

1540

events on transport of striped bass eggs to the estuarine turbidity maximum nursery area.

1541

Estuaries, 28, 108-123.

1542

56

1543

North, E. W., Hood, R. R., Chao, S. Y. and Sanford, L. P. (2006). Using a random displacement

1544

model to simulate turbulent particle motion in a baroclinic frontal zone: A new implementation

1545

scheme and model performance tests. Journal of Marine Systems, 60, 365-380.

1546 1547

Parada, C., Hinckley, S., Horne, J., Dorn, M., Hermann, A. and Megrey, B. A. (in press).

1548

Comparing simulated walleye pollock recruitment indices to data and stock assessment models

1549

from the Gulf of Alaska. Marine Ecology Progress Series.

1550 1551

Parada, C. (2003). Modeling the effects of environmental and ecological processes on the

1552

transport, mortality, growth and distribution of early life stages of Cape anchovy (Engraulis

1553

capensis) in the Bengula system. Ph.D. thesis, University of Cape Town, South Africa.

1554 1555

Parada, C., van der Lingen, C. D., Mullon, C. and Penven, P. (2003). Modelling the effect of

1556

buoyancy on the transport of anchovy (Engraulis capensis) eggs from spawning to nursery

1557

grounds in the southern Benguela: an IBM approach. Fisheries Oceanography, 12, 170-184.

1558 1559

Pedersen, O. P., Nilssen, E. M., Jorgensen, L. L. and Slagstad, D. (2006). Advection of the red

1560

king crab larvae on the coast of North Norway - A Lagrangian model study. Fisheries Research,

1561

79, 325-336.

1562 1563

Peliz, A., Marchesiello, P., Dubert, J., Marta-Almeida, M., Roy, C. and Queiroga, H. (in press). A

1564

study of crab larvae dispersal on the Western Iberian Shelf: physical processes. Journal of Marine

1565

Systems.

1566 1567

Penven, P. (2000). A numerical study of the Southern Benguela circulation with an application to

1568

fish recruitment. Ph.D. thesis, Université de Bretagne Occidentale, France.

1569 1570

Penven, P., Roy, C., Brundrit, G. B., de Verdière, A. C., Fréon, P., Johnson, A. S., Lutjeharms, J.

1571

R. E. and Shillington, F. A. (2001). A regional hydrodynamic model of upwelling in the Southern

1572

Benguela. South African Journal of Science, 97, 472-475.

1573

57

1574

Penven, P., Echevin, V., Pasapera, J., Colas, F. and Tam, J. (2005). Average circulation, seasonal

1575

cycle, and mesoscale dynamics of the Peru Current System: A modeling approach. Journal of

1576

Geophysical Research-Oceans, 110, C10021, doi:10010.11029/12005JC002945.

1577 1578

Penven, P., Debreu, L., Marchesiello, P. and McWilliams, J. C. (2006a). Evaluation and

1579

application of the ROMS 1-way embedding procedure to the central california upwelling system.

1580

Ocean Modelling, 12, 157-187.

1581 1582

Penven, P., Lutjeharms, J. R. E. and Florenchie, P. (2006b). Madagascar: a pacemaker for the

1583

Agulhas Current system? Geophysical Research Letters, 33, L17609,

1584

doi:17610.11029/12006GL026854.

1585 1586

Planque, B., Bellier, E. and Lazure, P. (2007). Modelling potential spawning habitat of sardine

1587

(Sardina pilchardus) and anchovy (Engraulis encrasicolus) in the Bay of Biscay. Fisheries

1588

Oceanography, 16, 16-30.

1589 1590

Rahel, F. J. (2002). Using current biogeographic limits to predict fish distributions following

1591

climate change. In Fisheries in a Changing Climate, ed. N. A. McGinn. American Fisheries

1592

Society Symposium 32, Bethesda, Maryland, pp. 99-110.

1593 1594

Railsback, S. F., Lamberson, R. H., Harvey, B. C. and Duffy, W. E. (1999). Movement rules for

1595

individual-based models of stream fish. Ecological Modelling 123, 73-89.

1596 1597

Rice, J. A., Quinlan, J. A., Nixon, S. W., Hettler Jr, W. F., Warlen, S. M. and Stegmann, P. M.

1598

(1999). Spawning and transport dynamics of Atlantic menhaden: inferences from characteristics

1599

of immigrating larvae and predictions of a hydrodynamic model. Fisheries Oceanography, 8, 93-

1600

110.

1601 1602

Robinson, C. L. K. and Ware, D. M. (1999). Simulated and observed response of the southwest

1603

Vancouver Island pelagic ecosystem to oceanic conditions in the 1990's. Canadian Journal of

1604

Fisheries and Aquatic Sciences, 56, 2433-2443.

58

1605 1606

Rodionov, S. and Overland, J. E. (2005). Application of a sequential regime shift detection

1607

method to the Bering Sea. ICES Journal of Marine Science, 62, 328-332.

1608 1609

Rodríguez, J. M., Barton, E. D., Eve, L. and Hernández-León, S. (2001). Mesozooplankton and

1610

ichthyoplankton distribution around Gran Canaria, an oceanic island in the NE Atlantic. Deep-

1611

Sea Research Part I-Oceanographic Research Papers, 48, 2161-2183.

1612 1613

Rose, K. A. and Sable, S. E. (2007). Multispecies modeling of fish populations. In Computers in

1614

fisheries research, Volume 2, eds. E. Moksness and B. A. Megrey. Chapman and Hall, New

1615

York, New York, pp. xxx-yyy.

1616 1617

Rose, K. A., Werner, F. E., Megrey, B. A., Aita, M. K., Yamanaka, Y, Hay, D. E., Schweigert, J.

1618

F. and Foster, M. B. (2007). Simulated herring growth responses in the Northeastern Pacific to

1619

historic temperature and zooplankton conditions generated by the 3-dimensional NEMURO

1620

nutrient-phytoplankton-zooplankton model. Ecological Modelling, 202, 184-195.

1621 1622

Rose, K. A., Megrey, B. A,. Hay, D. E., Werner, F. E., Schweigert, J. F. (in press). Climate

1623

regime effects on Pacific herring growth using coupled nutrient-phytoplankton-zooplankton and

1624

bioenergetics models. Transactions of the American Fisheries Society.

1625 1626

Rubec, P. J., Smith, S. G., Coyne, M. S., White, M., Sullivan, A., MacDonald, T. C., McMichael,

1627

R. H., Wilder, D. T., Monaco, M. E. and Ault, J. S. (2001). Spatial modeling of fish habitat

1628

suitability in Florida estuaries. In Spatial Processes and Management of Marine Populations, eds.

1629

G. H. Kruse, N. Bez, A. Booth, M. W. Dorn, S. Hills, R. N. Lipcius, D. Pelletier, C. Roy, S. J.

1630

Smith and D. Witherell. University of Alaska Sea Grant, AK-SG-01-02, Fairbanks, pp. 1-18.

1631 1632

Runge, J. A., Franks, P .J. S., Gentleman, W. C., Megrey, B. A., Rose, K. A., Werner, F. E.,

1633

Zakardjian, B. (2005). Diagnosis and prediction of variability in secondary production and fish

1634

recruitment processes: developments in physical-biological modelling. In: The Sea, Vol. 13, The

59

1635

global coastal ocean: multiscale interdisciplinary processes, eds. A. R. Robinson, K. H. Brink.

1636

Harvard University Press, pp. 413-473.

1637 1638

Russell, J. L., Stouffer, R. J. and Dixon, K. W. (2006). Intercomparison of the Southern Ocean

1639

circulations in IPCC coupled model control simulations. Journal of Climate, 19, 4560-4575.

1640 1641

Sætre, R., Toresen, R., Søiland, H. and Fossum, P. (2002). The Norwegian spring-spawning

1642

herring - spawning, larval drift and larval retention. Sarsia, 87, 167-178.

1643 1644

Santos, A. M. P., Peliz, A., Dubert, J., Oliveira, P. B., Angélico, M. M. and Ré, P. (2004). Impact

1645

of a winter upwelling event on the distribution and transport of sardine (Sardina pilchardus) eggs

1646

and larvae off western Iberia: a retention mechanism. Continental Shelf Research, 24, 149-165.

1647 1648

Santos, A. J. P., Nogueira, J. and Martins, H. (2005). Survival of sardine larvae off the Atlantic

1649

Portuguese coast: a preliminary numerical study. Ices Journal of Marine Science, 62, 634-644.

1650 1651

Santos, A. M. P., Ré, P., Dos Santos, A. and Peliz, A. (2006). Vertical distribution of the

1652

European sardine (Sardina pilchardus) larvae and its implications for their survival. Journal of

1653

Plankton Research, 28, 523-532.

1654 1655

Santos, A. M. P., Chícharo, A., Dos Santos, A., Moita, T., Oliveira, P. B., Peliz, A. and Ré, P.

1656

(2007). Physical-biological interactions in the life history of small pelagic fish in the Western

1657

Iberia Upwelling Ecosystem. Progress in Oceanography, 74, 192-209.

1658 1659

Scheffer, M., Baveco, J. M., DeAngelis, D. L., Rose, K. A. and van Nes, E. H. (1995). Super-

1660

individuals a simple solution for modelling large populations on an individual basis. Ecological

1661

Modelling, 80, 161-170.

1662 1663

Shannon, L. J., Moloney, C. L., Jarre, A. and Field, J. G. (2003). Trophic flows in the southern

1664

Benguela during the 1980's and 1990's. Journal of Marine Systems, 39, 83-116.

1665

60

1666

Shannon, L. J., Christensen, V. and Walters, C. J. (2004). Modelling stock dynamics in the

1667

southern Benguela ecosystem for the period 1978-2002. African Journal of Marine Science, 26,

1668

179-196.

1669 1670

Shchepetkin, A. F. and McWilliams, J. C. (2005). The regional oceanic modeling system

1671

(ROMS): a split-explicit, free-surface, topography-following-coordinate oceanic model. Ocean

1672

Modelling, 9, 347-404.

1673 1674

Schmitz, O. J., Post, E., Burns, C. E. and Johnston, K. M. (2003). Ecosystem responses to global

1675

climate change: moving beyond color mapping. BioScience, 53, 1199-1205.

1676 1677

Shin,Y. J. and Cury P. (2004). Using an individual-based model of fish assemblages to study the

1678

response of size spectra to changes in fishing. Canadian Journal of Fisheries and Aquatic

1679

Sciences, 61, 414-431.

1680 1681

Shin Y. J., Shannon, L. J. and Cury, P.M. (2004). Simulations of fishing effects on the southern

1682

Benguela fish community using an individual-based model: learning from a comparison with

1683

ECOSIM. African Journal of Marine Science, 26, 95-114.

1684 1685

Skogen, M. (1999). A biophysical model applied to the Benguela upwelling system. South

1686

African Journal of Marine Science, 21, 235-249.

1687 1688

Skogen, M. D., Shannon, L. J. and Stiansen, J. E. (2003). Drift patterns of anchovy Engraulis

1689

capensis larvae in the southern Benguela, and their possible importance for recruitment. African

1690

Journal of Marine Science, 25, 37-47.

1691 1692

Stenevik, E. K., Skogen, M., Sundby, S. and Boyer, D. (2003). The effect of vertical and

1693

horizontal distribution on retention of sardine (Sardinops sagax) larvae in the Northern Benguela

1694

- observations and modelling. Fisheries Oceanography, 12, 185-200.

1695

61

1696

Suda, M. and Kishida, T. (2003). A spatial model of population dynamics of the early life stages

1697

of Japanese sardine, Sardinops melanostictus, off the Pacific coast of Japan. Fisheries

1698

Oceanography, 12, 85-99.

1699 1700

Suda, M., Akamine, T. and Kishida, T. (2005). Influence of environment factors, interspecific-

1701

relationships and fishing mortality on the stock fluctuation of the Japanese sardine, Sardinops

1702

melanostictus, off the Pacific coast of Japan. Fisheries Research, 76, 368-378.

1703 1704

Sugden, A. and Pennisi, E. (2006). When to go, where to stop. Science, 313, 775.

1705 1706

Thorpe, S. E., Heywood, K. J., Stevens, D. P. and Brandon, M. A. (2004). Tracking passive

1707

drifters in a high resolution ocean model: implications for interannual variability of larval krill

1708

transport to South Georgia. Deep-Sea Research Part I-Oceanographic Research Papers, 51, 909-

1709

920.

1710 1711

Tilburg, C. E., Reager, J. T. and Whitney, M. M. (2005). The physics of blue crab larval

1712

recruitment in Delaware Bay: A model study. Journal of Marine Research, 63, 471-495.

1713 1714

Twatwa, N. M., van der Lingen, C. D., Drapeau, L., Moloney, C. L. and Field, J. G. (2005).

1715

Characterising and comparing the spawning habitats of anchovy Engraulis encrasicolus and

1716

sardine Sardinops sagax in the southern Benguela upwelling ecosystem. African Journal of

1717

Marine Science, 27, 487-500.

1718 1719

Tyler, J. A. and Rose, K. A. (1994). Individual variability and spatial heterogeneity in fish

1720

population models. Reviews in Fish Biology and Fisheries, 4, 91-123.

1721 1722

van der Lingen, C. D., Castro, L., Drapeau, L. and Checkley, D. (2005). Report of a GLOBEC-

1723

SPACC workshop on characterizing and comparing the spawning habitats of small pelagic fish.

1724

GLOBEC Report 21: xii.

1725

62

1726

Vikebø, F., Sundby, S., Ådlandsvik, B. and Fiksen, Ø. (2005). The combined effect of transport

1727

and temperature on distribution and growth of larvae and pelagic juveniles of Arcto-Norwegian

1728

cod. Ices Journal of Marine Science, 62, 1375-1386.

1729 1730

Vikebø, F. B., Sundby, S., Ådlandsvik, B. and Otterå, O. H. (2007). Impacts of a reduced

1731

thermohaline circulation on transport and growth of larvae and pelagic juveniles of Arcto-

1732

Norwegian cod (Gadus morhua). Fisheries Oceanography, 16, 216-228.

1733 1734

Vikebø, F., Jørgensen, C., Kristiansen, T. and Fiksen, O. (in press). Drift, growth, and survival of

1735

larval Northeast Arctic cod with simple rules of behaviour. Marine Ecology-Progress Series.

1736 1737

Walters C., Christensen V. and Pauly, D. (1997). Structuring dynamic models of exploited

1738

ecosystems from trophic mass-balance assessments. Reviews in Fish Biology and Fisheries, 7,

1739

139-172.

1740 1741

Walters, C., Pauly D. and Christensen, V. (1999). Ecospace: prediction of mesoscale spatial

1742

patterns in trophic relationships of exploited ecosystems, with emphasis on the impacts of marine

1743

protected areas. Ecosystems, 2, 539-554.

1744 1745

Walters, C., Pauly D., Christensen, V. and Kitchell, J. F. (2000). Representing density dependent

1746

consequences of life history strategies in aquatic ecosystems: EcoSim II. Ecosystems, 3, 70-83.

1747 1748

Werner, F. E., Quinlan, J. A., Lough, R. G. and Lynch, D. R. (2001). Spatially-explicit individual

1749

based modeling of marine populations: a review of the advances in the 1990s. Sarsia, 86, 411-

1750

421.

1751 1752

Werner, F. E., Ito, S., Megrey, B. A. and Kishi, M. J. (2007). Synthesis of the NEMURO model

1753

studies and future directions of marine ecosystem modeling. Ecological Modelling, 202, 211-223.

1754 1755

Yeung, C. and Lee, T. N. (2002). Larval transport and retention of the spiny lobster, Panulirus

1756

argus, in the coastal zone of the Florida Keys, USA. Fisheries Oceanography, 11, 286-309.

63

1757 1758

Zakardjian, B. A., Sheng, J. Y., Runge, J. A., McLaren, I., Plourde, S., Thompson, K. R. and

1759

Gratton, Y. (2003). Effects of temperature and circulation on the population dynamics of Calanus

1760

finmarchicus in the Gulf of St. Lawrence and Scotian Shelf: Study with a coupled, three-

1761

dimensional hydrodynamic, stage-based life history model. Journal of Geophysical Research-

1762

Oceans, 108, 8016, doi:8010.1029/2002JC001410.

1763

64