Ecological Modelling 255 (2013) 38–57

Contents lists available at SciVerse ScienceDirect

Ecological Modelling

jo urnal homepage: www.elsevier.com/locate/ecolmodel

Trophodynamics of the eastern ecosystem: Ecological

change associated with the growth of ’s largest fishery

a,∗ a,b a,c d a,e,b

Simon D. Goldsworthy , Brad Page , Paul J. Rogers , Cathy Bulman , Annelise Wiebkin ,

a,e,b a,e a,e,f a g

Lachlan J. McLeay , Luke Einoder , Alastair M.M. Baylis , Michelle Braley , Robin Caines ,

e a,c a,e a a

Keryn Daly , Charlie Huveneers , Kristian Peters , Andrew D. Lowther , Tim M. Ward

a

South Australian Research and Development Institute, Aquatic Sciences, 2 Hamra Ave, West Beach, 5024, Australia

b

Department of Environment, Water and Natural Resources, GPO 1047, Adelaide, South Australia 5001, Australia

c

School of Biological Sciences, Flinders University, GPO Box 2100, Adelaide, South Australia 5001, Australia

d

CSIRO Marine Research, GPO Box 1538, Hobart, Tasmania 7001, Australia

e

University of Adelaide, School of Earth and Environmental Sciences, North Terrace, Adelaide, South Australia 5005, Australia

f

Deakin University, Princess Highway, Warrnambool, Victoria 3280, Australia

g

University of South Australia, Mawson Lakes, South Australia 5095, Australia

a r t i c l e i n f o a b s t r a c t

Article history: We used the Ecopath with Ecosim software to develop a trophic mass-balance model of the eastern

Received 17 May 2012

Great Australian Bight ecosystem, off southern Australia. Results provide an ecosystem perspective of

Received in revised form 21 January 2013

Australia’s largest fishery, the South Australian sardine fishery, by placing its establishment and growth

Accepted 23 January 2013

in the context of other dynamic changes in the ecosystem, including: the development of other fish-

Available online 1 March 2013

eries; changing abundances of apex predator populations and oceanographic change. We investigated

the potential impacts of the sardine fishery on high tropic level predators, particularly land-breeding

Keywords:

seals and seabirds which may be suitable ecological performance indicators of ecosystem health. Results

Eastern Great Australian Bight

indicate that despite the rapid growth of the sardine fishery since 1991, there has likely been a negli-

Food web model

gible fishery impact on other modelled groups, suggesting that current levels of fishing effort are not

Ecopath with Ecosim

Sardine impacting negatively on the broader ecosystem structure and function in the eastern Great Australian

Sardinops sagax Bight. Results highlight the importance of small pelagic fish to higher trophic levels, the trophic changes

Fishing impacts that have resulted from loss and recovery of apex predator populations, and the potential pivotal role

of cephalopod biomass in regulating ‘bottom-up’ trophic processes. The ability to resolve and attribute

potential impacts from multiple fisheries, other human impacts and ecological change in this poorly

understood region is highlighted by the study, and will be critical to ensure future ecologically sustainable

development within the region.

Crown Copyright © 2013 Published by Elsevier B.V. All rights reserved.

1. Introduction as important in marine ecosystems in the transfer of production

from plankton to higher trophic level species (marine mammals,

There is growing concern about the global impacts of increased seabirds and large predatory fish), and concerns about the poten-

fishing effort on ecosystem structure and function. Much of this tial impacts on high trophic level species from localised depletion

concern stems from the growth of fisheries that target low trophic by fisheries has been identified by many studies (Cury et al., 2000,

level species (planktivorous small pelagic fish), which currently 2011; Jennings et al., 2012; Smith et al., 2011). Although some

account for 30% of global landings (Ainley and Blight, 2009; studies have successfully demonstrated the relationships between

Branch et al., 2010). This growth has been largely driven by overfishing of low trophic level species and the impacts on higher

increased demand for fish meal production, feed or fertiliser for trophic levels, especially seabirds (Cury et al., 2011; Frederiksen

the aquaculture and agriculture industries and only a small portion et al., 2005; Jahncke et al., 2004; Myers et al., 2007), the numerical

is processed for human consumption (Alder et al., 2008; Merino and trophodynamic relationships between predators and prey are

et al., 2010; Naylor et al., 2009). Small pelagic fishes are recognised poorly understood, even in regions where there are commercially

valuable fisheries (Abrams and Ginzburg, 2000; Cury et al., 2011). As

such, numerical modelling of ecosystem dynamics is increasingly

∗ being utilised to resolve these complex relationships (see review by

Corresponding author.

E-mail address: [email protected] (S.D. Goldsworthy). Fulton, 2010) and to achieve ecosystem based fishery management

0304-3800/$ – see front matter. Crown Copyright © 2013 Published by Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.ecolmodel.2013.01.006

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39

(EBFM) objectives (Pikitch et al., 2004). The latter is dependent The EGAB supports high densities of small pelagic fishes,

on the development of reliable indicators of ecosystem status and the dominant species being sardine and anchovy (Engraulis

health (Cury et al., 2011; Lozano-Montes et al., 2011; Pikitch et al., australis) (Ward et al., 2006). Other important small pelagic

2004; Samhouri et al., 2009). The objectives underpin ecologically fish species include blue mackerel (Scomber australasicus), jack

sustainable development (ESD) frameworks that form the basis for mackerel (Trachurus declivis), yellowtail scad (T. novaezelandiae),

modern management of Australian fisheries; however, there are redbait (Emmelichthys nitidus), maray (Etrumeus teres) and saury

few examples where they are achieved or put into practice. They (Scomberesox saurus). Primary and secondary production in the

include: (i) avoid degradation of ecosystems, (ii) minimise risk of EGAB appears to be underpinned by a unique northern bound-

irreversible change to assemblages of species and ecosystem pro- ary current system, the Flinders Current (Middleton and Bye, 2007;

cesses, (iii) obtain and maintain long-term socioeconomic benefits Middleton and Cirano, 2002; van Ruth et al., 2010a, 2010b; Ward

without compromising the ecosystem, and (iv) generate knowledge et al., 2006), the dominant oceanographic feature of which is coastal

of ecosystem processes sufficient to understand the consequence upwelling that occurs between November to May, especially off

of human actions (Pikitch et al., 2004). the Bonney Coast, and southern and western Eyre

The South Australian sardine (Sardinops sagax) fishery (sar- Peninsula (McClatchie et al., 2006; Middleton and Bye, 2007; van

dine fishery) is Australia’s largest fishery by mass of catch. It was Ruth et al., 2010a, 2010b; Ward et al., 2006) (Fig. 1). This coastal

established in 1991 to provide feed for the southern bluefin tuna upwelling has oceanographic, biological and ecological similari-

(SBT, Thunnus maccoyii) mariculture industry in Port Lincoln, South ties to the larger and more productive eastern boundary currents

Australia (Fig. 1) (Ward et al., 2005). Operated as a purse seine fish- that underpin the Benguela, Humbolt, California and Canary Cur-

ery, sardine catch within this region is centred on southern Spencer rent upwelling systems (Mann and Lazier, 1996; Middleton and

Gulf, Investigator Strait and the western Eyre Peninsula in the east- Platov, 2003; Ward et al., 2006), although the latter typically have

ern Great Australian Bight (EGAB) (Ward et al., 2005). Spawning an order of magnitude higher productivity and small pelagic fish

biomass of sardine was estimated to be 165,000 t in 1995 (Ward biomass compared to the EGAB (Guénette et al., 2008).

et al., 2009), but it declined by over 70% to ∼37,000 t in 1996 The rich pelagic resources of the EGAB underpin arguably the

following an unprecedented mass mortality event, recovered to greatest density and biomass of apex predators to be found in

∼146,000 t in 1998 and then declined by over 70% again to ∼36,000 t Australian coastal waters. These include marine mammals such as

in early 1999 following a second mass mortality event (Ward et al., pygmy blue whales (Balaenoptera musculus brevicauda), and >80%

2001). Between 1994 and 2001, fishery catches ranged between of Australia’s populations of New Zealand fur seals (Arctocephalus

2500 and 6500 t each year, but the total allowable commercial forsteri) and Australian sea lions (Neophoca cinerea) (Branch et al.,

catch (TACC) was expanded and the harvest increased markedly 2007; Goldsworthy et al., 2003, 2009), and the Australian fur seal (A.

to 42,475 t in 2005, and then stabilising around 30,000 t from 2007. pusillus doriferus) (Shaughnessy et al., 2010). All seal species were

Fig. 1. Location of the eastern Great Australian Bight (EGAB) ecosystem included in the Ecopath with Ecosim model. The depth contours are 100, 200, 500, 1000 and 2000 m.

40 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57

subjected to early colonial sealing, with recovery of fur seal popula- of the EGAB ecosystem. Ecopath was developed by Polovina

tions commencing in the 1970s and 1980s (Goldsworthy and Page, (1984), based on a simple steady-state trophic box model, and

2007; Ling, 1999; Shaughnessy and McKeown, 2002). Other key further developed by Christensen and Pauly (1992) and Walters

apex predators include seabirds, such as short-tailed shearwaters et al. (1997). Whereas Ecopath enables description of the static

(Puffinus tenuirostris), little penguins (Eudyptula minor) and crested state energy flow of an ecosystem at a particular point in time,

terns (Sterna bergii); pelagic sharks including bronze whalers (Car- Ecosim enables dynamic simulations based on Ecopath parameters

charhinus brachyurus) and dusky shark (C. obscurus), white shark that allow the estimation of ecosystem response to environmental

(Carcharodon carcharias) and shortfin mako (Isurus oxyrinchus); perturbations. The Ecopath with Ecosim software has now been

and predatory fishes, such as southern bluefin tuna and yellowtail used to describe a diverse range of aquatic ecosystems world-wide,

kingfish (Seriola lalandi) (Copley, 1996; Goldsworthy et al., 2003; and details of the ecological theory and mathematical equations

Goldsworthy and Page, 2007; Ward et al., 2006). that underpin its key functions have been extensively detailed

The EGAB supports some of Australia’s most valuable fisheries, elsewhere (e.g. Christensen and Walters, 2004; Griffiths et al.,

including four main Commonwealth and five main South Australian 2010; Piroddi et al., 2010; Shannon et al., 2008).

(State) managed fisheries (Knight and Tsolos, 2010; Wilson et al.,

2009). The main Commonwealth fisheries that operate are the 2.2. Model area and structure

southern bluefin tuna purse seine fishery (SBT), GAB Trawl (GABT),

South East Trawl (SET), and shark gillnet component of the Gill Hook We modelled the eastern Great Australian Bight (EGAB) pelagic

and Trap fishery (GHAT) (Wilson et al., 2009). The main South Aus- ecosystem, a region off the South Australian coast that includes con-

◦ ◦

tralian State managed fisheries that operate in the region are the tinental shelf waters to 200 m depth between 132 E and 139.7 E.

sardine fishery, southern rock lobster, abalone, western king prawn The model domain includes Investigator Strait and the mid-lower

and marine scalefish (MSF) fisheries (Knight and Tsolos, 2010). portions of and where the sardine fish-

∼ 2

In recognition of the economic importance of the sardine fishery, ery is centred. The total modelled area was 154,084 km (Fig. 1).

the important role of small pelagic fishes in underpinning ecologi- This area was selected because the distribution and abundance of

cal processes in the EGAB ecosystem, and the high socio-economic sardine and other small pelagic fish species was well known from

and conservation significance of the region’s marine predators, small pelagic fish ichthyoplankton surveys between 1995 and 2010

the fishing industry and fishery managers identified the need to (Ward et al., 2007, 2009).

manage potential ecological impacts. In response to this need, a As the aims of the model were to investigate the potential

large-scale ecosystem programme was initiated to assess the role impacts of the sardine fishery on high tropic level predators,

of sardines in the EGAB ecosystem and to develop ecological perfor- especially land-breeding seals and seabirds, these groups were dis-

mance indicators and reference points for the fishery. As part of this aggregated into single species where data on diet and population

programme, this study aimed to: (1) develop a food web model for biomass permitted. As we were also interested in the broad ecolog-

the EGAB and describe its temporal dynamics over the period since ical roles and importance of small pelagic fish in general, including

the sardine fishery established in 1991 using time-series data of sardines, sardines, anchovies, jack mackerel, redbait, and other key

fishing activity and environmental drivers; (2) identify the trophic mid-trophic level species such as arrow squid and calamary, were

interactions and functional groups most sensitive to variability in also examined as single-species functional groups where data per-

sardine biomass; (3) determine the sensitivity of land breeding mitted.

apex predators to changes in sardine biomass and fishery catch, and A total of 40 functional and species groups were devel-

assess their appropriateness as ecological performance indicators oped based on species similarity in terms of diet, habitat,

for the sardine fishery; (4) examine temporal changes to ecosystem foraging behaviour, size, consumption and rates of production

structure using ecosystem indicators and assess their potential as (Supplementary Information A and B). Intrinsic to Ecopath model

ecological performance indicators of ecosystem health; (5) exam- development, each trophic group operates as a single biomass,

ine the trophodynamic changes occurring in response to changes in despite groups often being composed of several species. The aggre-

apex predator biomasses, including recovering fur seal populations gation of species into trophic groups will therefore impact on model

and declining SBT populations and (6) examine how the recovery dynamics in some instances; however, by matching species for

of these groups is likely to impact on the EGAB ecosystem and the diet, consumption, and production rates we attempted to constrain

sardine fishery. the errors and uncertainty of aggregating. The model was para-

Despite global recognition of the need to adopt sustainable meterised for 1991, when the sardine fishery commenced. Model

approaches to the management of fisheries and aquaculture simulations were run over the following 18 years (1991–2008).

(Pikitch et al., 2004), there are still few examples where EBFM For each trophic group, four inputs were derived from the avail-

has been adopted to complement the existing single-species/stock able data: diet, biomass, production per unit of biomass (P/B) and

management paradigm. The principal aim of this study was consumption per unit of biomass (Q/B). A dietary matrix was con-

to provide an ecosystem perspective of the sardine fishery, by structed using empirical data derived within the EGAB ecosystem

placing its establishment and growth in the context of other where possible (Table 1). For marine mammals, seabirds, large and

dynamic changes in the ecosystem, including changes to other small pelagic fishes, arrow squid and calamary, dietary data summ-

fisheries, changes to apex predator populations, and meteorolog- arised by Page et al. (2011) was used (Table 1). For most other

ical and oceanographic changes. Such information is essential to fish species in the region, data from Currie and Sorokin (2010) was

−2

enhance EBFM of Australia’s largest pelagic fishery by incorporat- used. Where possible, the biomasses (t km ) of functional groups

ing scientifically-based approaches for assessing and, if necessary, were estimated either from field surveys or stock assessments. A

mitigating the fishery’s trophic impacts. detailed description of the functional groups and how estimates of

biomass, P/B and Q/B were derived is presented as Supplementary

2. Materials and methods Information (A and B).

Fishery data on landings, discards and effort between 1991 and

2.1. Ecopath and mass balance approach 2008 were obtained from State and Commonwealth logbook data.

A total of 11 fisheries (fleets) operate within the EGAB ecosystem

We used the Ecopath with Ecosim software v 6.1.0 (Fig. 2, Tables 2 and 3). Six were South Australian (State) managed

(www.Ecopath.org) to develop a trophic mass-balance model fisheries: the South Australian (SA) sardine, SA Marine Scalefish

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 41 0.0163 0.5799 0.1960 0.0073 0.1841 0.0163 19

0.0137 0.7438 0.1197 0.1164 0.0064 18

0.0076 0.7604 0.1834 0.0292 0.0194 17

0.0303 0.0360 0.0909 0.0375 0.2164 0.0003 0.0909 0.0502 0.3771 0.0147 0.0108 0.0234 0.0092 0.0118 0.0002 16

0.0777 0.0447 0.0356 0.0433 0.5267 0.0370 0.1341 0.0232 0.0777 15

0.0606 0.0720 0.1818 0.0750 0.4288 0.1818 14

0.0045 0.0009 0.2219 0.7727 13

0.1000 0.1187 0.0964 0.0237 0.2790 0.0524 0.2032 0.1264 12

0.0133 0.0064 0.0818 0.0002 0.0288 0.0019 0.0781 0.2298 0.0072 0.0021 0.0004 0.0001 0.1904 0.2166 0.0814 0.0603 0.0012 11

0.0009 0.0586 0.0043 0.0016 0.0141 0.3072 0.2294 0.0783 0.0249 0.0567 0.1612 0.0007 0.0007 0.0527 0.0021 0.0037 0.0002 0.0029 10

0.3850 0.1100 0.0900 0.1600 0.0075 0.0200 0.0375 0.1500 0.0200 0.0200 9 model.

Ecopath 0.0153 0.0002 0.0509 0.0045 0.0044 0.0004 0.0002 0.0161 0.0104 0.0271 0.0004 0.0098 0.0361 0.0082 0.8161 8

EGAB

0.0006 0.0161 0.0441 0.6152 0.0353 0.0406 0.0885 0.0053 0.0075 0.0001 0.1137 0.0273 0.0055 7

balanced

the

0.0004 0.0025 0.0042 0.0954 0.0083 0.0010 0.0025 0.1097 0.0075 0.0022 0.0161 0.0151 0.0475 0.1472 0.1905 0.3029 0.0468 6 in

used

0.0640 0.0536 0.4489 0.3674 0.0029 0.0027 0.0012 0.0048 0.0460 0.0034 0.0051 5

0.0015 0.0010 0.0712 0.0011 0.0611 0.1186 0.0055 0.0109 0.0003 0.0145 0.0630 0.0375 0.0021 0.0043 0.0301 0.0050 0.2697 0.0022 0.0008 0.0013 0.0004 0.2979 4 contribution

prey

0.0862 0.0032 0.1137 0.4328 0.2082 0.0063 0.0170 0.0208 0.0019 0.0095 0.0124 0.0602 0.0248 0.0029 3

proportional

0.0406 0.0004 0.0153 0.0209 0.0013 0.0087 0.2326 0.0140 0.0467 0.0007 0.1337 0.2004 0.2846 2

showing

0.0250 0.4750 0.5000 1

matrix

1 2 3 4 5 6 7 8 9 10 20 30 40 22 36 27 24 37 16 34 29 35 Group 11 12 13 14 15 17 18 19 23 25 28 31 32 33 38 39 41 26 diet

their

-

and

piscs invert

ruffs

(infauna dolphin

sharks dolphin

skates

groups

and demersal demersal

lion

feeders

production seal

grazer small

sharks

whales

squid seal

name

tunas-kingfish squids

demersal demersal demersal zooplankton bentho-pelagic zooplankton feeders

penguin

and

fur sea mackerel

mackerel

1

fur

planktivores (megabenthos) feeders piscs invert macrobenthos) piscs (carnivores) omnivore (herbivores) Group Baleen Gannets Pelagic Import Demersal Large Bottlenose Common NZ Aust Aust Little Terns Petrels Rays SBT Other Small Medium Arrow Small Salmons Jack Redbait Anchovy SardineInshore 21 Mesopelagics Calamary Octopus Large Small Other Blue Medium Small Benthic Detritivore Filter Primary Detritus Table Functional

42 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 0.8000 0.2000 38

0.1000 0.4000 37

0.4000 0.5500 36

35 1.0000

0.2000 0.8000 34

0.0807 0.0050 0.0538 0.0538 0.6992 0.1076 33

0.0500 0.0600 0.0020 0.0080 0.8800 32

0.0990 0.3940 0.1120 0.1050 0.0140 0.1430 0.0270 0.0050 0.0160 0.0850 31

0.1032 0.0400 0.0016 0.0744 0.0371 0.1213 0.0075 0.0088 0.0031 0.0109 0.2202 0.0112 0.0095 0.0381 0.2626 0.0086 0.0407 0.0012 30

0.0001 0.1684 0.0024 0.6602 0.0479 0.0106 0.1104 29

0.0020 0.0600 0.0080 0.9300 28

0.0377 0.0118 0.9163 0.0342 27

0.0300 0.0427 0.1310 0.0040 0.6292 0.1574 0.0018 0.0039 26

0.0033 0.0003 0.1129 0.0347 0.1036 0.0815 0.0710 0.5926 25

0.0035 0.0150 0.0091 0.0115 0.0693 0.0056 0.0825 0.1864 0.4419 0.1752 24

0.0420 0.1920 0.0980 0.3251 0.0245 0.0442 0.2205 0.0245 23

0.8652 0.1348 22

0.0296 0.6748 0.2883 0.0049 0.0023 21

0.4938 0.5062 20

2 3 5 9 30 40 22 36 27 0.0093 24 37 16 34 2935 0.0200 Group 14 17 21 23 25 28 31 33 38 39 41 26

-

) piscs invert

ruffs

(infauna dolphin

sharks 12 dolphin

skates 13

and demersal demersal

lion 6 feeders production seal

grazer small

sharks 11

whales 1

squid Continued seal 4 name

tunas-kingfish 15 squids 32

demersal demersal demersal zooplankton ( bentho-pelagic zooplankton feeders

penguin 7

and

fur sea mackerel

mackerel 18

1

fur

planktivores (megabenthos) feeders piscs invert macrobenthos) piscs (carnivores) omnivore (herbivores) Rays Large Aust Bottlenose Group Baleen Demersal Little SBT PetrelsGannets 8 Other Import Common NZ Aust TernsPelagic 10 Jack AnchovyInshore Salmons 20 Calamary Other Medium Small Mesopelagics Octopus Detritivore Primary Small Benthic Detritus Blue RedbaitSardine Medium 19 Small Arrow Large Small Filter Table

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 43

catch and catch per unit-effort (CPUE) were estimated for 18 and

16 functional groups, respectively. A time series of annual fish-

ing mortality (F) was used for sardines instead of CPUE because

marked changes in vessel size in the fleet between 1991 and 2008

confounded CPUE estimates (Supplementary Information C).

There were few data on protected species bycatch in most

fisheries; however, estimates were available for common dolphin

and bycatch in the sardine fishery and demersal

gillnet shark fishery, respectively (Goldsworthy et al., 2010; Hamer

et al., 2008). Sea lion bycatch was included as discards, as the

fishery overlaps with the foraging distribution of sea lions across

the model region (bycatch scaled to 2008 estimated bycatch

relative to historic effort); estimates of dolphin bycatch in the

sardine fishery were available between 2005 and 2008 (based

on bycatch rates per unit of fishing effort) (Hamer et al., 2008),

and were estimated for the period between 1991 and 2004 using

fishing effort. The sardine fishery impacts dolphins only in part of

their range, hence bycatch was estimated as fishing mortality (F)

by running the model without bycatch mortality, then estimating

F directly from the reconstructed bycatch time-series. The model

was then re-run including F. Some common dolphin bycatch is

also known to occur in the demersal gillnet shark fishery, but no

quantitative data were available.

The mixed trophic impacts routine in the network analysis tools

within Ecopath was used to evaluate critical trophic interactions

between groups in the ecosystem (Supplementary Information, Fig.

S1). The routine is based on the method developed by Ulanowicz

and Puccia (1990), and allows the computation of direct and

indirect impacts which a change in biomass of a predator group

will have on other groups in the system, assuming that the diet

matrix remains unchanged, and may thus be viewed as a tool for

sensitivity analysis.

2.3. Model fitting

Model fitting using the ‘fit to time series’ procedure was

run in Ecosim using the time-series estimates (1991–2008) of

biomass and fishing mortality (F) (sardines only), catch and CPUE

for functional groups with available data. Different numbers of

predator–prey interactions within the dietary matrix were selected

(10–90) within this procedure to identify the most sensitive and

optimal number of predator–prey interactions, and their vulner-

ability values that would minimise the model sum of squares (SS)

−1

Fig. 2. Trends in the total landings (catch t y ) from eight main fisheries in the and produce the best fit to the time series data (a table with the esti-

EGAB ecosystem between 1991 and 2008. (A) All fisheries combined (Total catch),

mated vulnerabilities is provided in Supplementary Information D).

the SASF (Sardine catch) and all fisheries minus sardine catch (Other catch); (B)

Some of the default Ecosim parameters were then adjusted to

marine-scalefish line (MS-line); marine-scalefish net (MS-net); Spencer Gulf, Gulf

further decrease the model SS. This included adjusting the max-

St Vincent, West Coast prawn fisheries (Prawn); demersal shark (Shark gillnet); GAB

trawl fishery (GABT); South-east Trawl fishery (SET); and Southern bluefin tuna pole imum relative feeding time of marine mammals and seabirds

and line and purse seine fishery (SBT).

from 2.0 (default) to 10.0, and their feeding time adjustment

rates to 0.5 (0 for all other groups), to account for modifica-

line, SA Marine Scalefish net, and three prawn fisheries (Spencer tions to their search feeding times in response to changes in

Gulf; Gulf St Vincent; West Coast). Another five fisheries are man- prey availability (Christensen et al., 2008). Similarly, we adjusted

aged by the Australian Government (Commonwealth): Southern density-dependent predator–prey switching power of the dolphin

Bluefin Tuna (SBT) purse seine, SBT pole and bait, South East trawl, and seal groups from 0 to 2.0, to account for their capacity to

Great Australian Bight (GAB) trawl, and the gillnet demersal shark opportunistically adjust their diet in response to changes in prey

fishery (part of the Gillnet Hook and Trap Fishery). Fleet data availability (Piroddi et al., 2010). We also explored improvements

from the SA abalone and southern rock lobster fisheries were not to model fits by adjusting values of density-dependent changes in

included because the EGAB ecosystem model was primarily devel- catchability for pelagic schooling fish such as sardines and tuna

oped to be a pelagic model. (Christensen et al., 2008; Piroddi et al., 2010), but these did not

Retained and discarded catch (landings and discards, respec- produce improvements to model fitting.

tively) data were typically only available for between 1 and 3 years The final step of the model fitting procedure was to link the

for each fishery, and were estimated for 1991 based on their propor- best combination of vulnerability values and the time-series data

tion to landed catch (Currie et al., 2009; Fowler et al., 2009; Roberts to an estimated trend in primary productivity (PP) within the

and Steer, 2010). All landed and discarded species were assigned EGAB ecosystem. The importance of seasonal upwelling in the

their functional group, and summed (catch) at the functional group EGAB to enhancing primary, secondary and fish production has

−2

level (t km ) (Tables 2 and 3, respectively). Time series of annual been established (Ward et al., 2006). As such, time-series of mean

44 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 3 2 4 2 2 3 7 5 3 3 6 5 5 3 8 3 2 6 7 7 − − − − − − − − − − − − − − − − − − − − 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10

× × × × × × × × × × × × × × × × × × × ×

Total 1.62 2.92 2.16 1.38 3.42 1.95 1.09 5.72 4.78 6.21 8.47 1.18 6.00 6.27 1.87 3.25 1.00 2.16 7.87 10 9 9 − − − 10 10 10

× × ×

prawn

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1.48 2.11 2.26 fishery WC 7 6 6 − − − 10 10 10

prawn × × ×

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 GSV fishery 1.26 1.80 1.93 4 3 3 − − − 10 10 10

× × ×

prawn

00 0 0 0 0 00 0 0 0 00 0 0 0 8.22 0 0 0 fishery 1.51 2.15 2.30 SG 5 3 4 4 4 3 6 7 7 7 7 7 − − − − − − − − − − − − 10 10 10 10 10 10 10 10 10 10 10 10

× × × × × × × × × × × ×

0 0 0 0 0 0 0 0 1.54 6.28 2.42 2.51 1.00 5.00 1.83 3.30 6.00 1.00 3.00 6.98 shark 5 6 7 4 7 8 8 8 6 3 3 5 − − − − − − − − − − − − fishery.

10 10 10 10 10 10 10 10 10 10 10 10

trawl Demersal × × × × × × × × × × × ×

0 0 0 0 0 0 0 2.26 2.23 1.95 3.36 1.95 6.49 1.05 1.09 3.25 3.25 1.45 1.42 sardine

5 5 3 4 6 6 5 3 − − − − − − − − 10 10 10 10 10 10 10 10

Australian × × × × × × × ×

trawl GAB

0 0 0 0 0 0 0 0 0 0 0 2.16 1.38 5.95 8.47 8.47 5.18 1.97 SE South

=

6 5 6 4 − − − − SASF

10 10 10 10

× × × ×

bait pole

model.

0 0 0 0 0 0 0 0 0 0 0 0 0 0 9.23 9.17 1.03 1.02 and SBT 2 2 2 − − − Ecopath

10 10 10

× × ×

purse

EGAB

0000 0 0 00 0 0 0 00 0 0 1.27 0 0 0 0 0 0 SBT 2.14 1.35 3.48 seine 3 3 4 3 3 4 2 5 balanced

− − − − − − − − 10 10 10 10 10 10 10 10

the

× × × × × × × ×

in

0 0 0 0 0 0 0 0 0 0 0 SAMS-net fisheries 8.11 8.21 6.00 4.47 4.69 1.73 7.49 2.40 used

5 3 4 3 3 4 5 3 − − − − − − − − group

10 10 10 10 10 10 10 10

× × × × × × × ×

0 0 0 0 0 0 0 0 0 fisheries 1.66 2.29 9.05 9.66 7.06 functional

by

5 5 ) − − 2 − 10 10

× × km

t

0 0 0 0 0 0 0 0 0 000 0 0 2.65 0 1.37 5.72 (catch

piscs invert landings

ruffs

fleet

sharks

skates

and demersal demersal

feeders

of

grazer

sharks 0 9.13

squid 0 name SASF SAMS-line

tunas-kingfish squids

demersal demersal bentho-pelagic

and mackerel 0

mackerel 0 0

2

(megabenthos) feeders piscs invert piscs Group Pelagic Demersal Rays SBTOther Large Blue Jack SardineSalmons Medium 0Small Medium Small 5.72 0 Arrow Calamary Other Octopus Benthic Sum Table Summary

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 45 5 5 4 3 7 4 7 7 5 3 5 4 3 4 4 2 3 6 6 7 3 7 3 − − − − − − − − − − − − − − − − − − − − − − − 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10

× × × × × × × × × × × × × × × × × × × × × × ×

2.17 4.49 7.57 1.08 8.40 2.80 2.56 2.22 2.58 2.88 3.79 3.09 4.85 2.08 3.12 1.94 1.99 1.98 4.10 4.69 2.20 1.11 1.51 Total 10 10 13 13 14 10 10 10 10 10 9 10 9 13 9 9 11 − − − − − − − − − − − − − − − − − 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10

× × × × × × × × × × × × × × × × ×

prawn

0 0 0 0 0 0 fishery 6.33 9.04 1.05 1.84 2.82 3.71 3.02 6.51 6.94 3.04 1.71 1.95 3.57 4.59 2.04 9.80 9.82 WC 7 − 7 7 7 10 10 10 11 7 7 7 6 8 6 7 6 7 − − − − − − − − − − − − − − − − 10

10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10

×

prawn × × × × × × × × × × × × × × × ×

0 0 0 0 0 0 GSV 5.40 7.72 8.96 1.57 2.41 3.17 2.58 5.56 5.93 2.60 1.461 1.66 3.05 3.92 1.74 8.37 8.39 fishery 4 4 4 7 7 8 3 4 3 4 3 7 4 3 3 4 2 − − − − − − − − − − − − − − − − − 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10

× × × × × × × × × × × × × × × × ×

prawn

0 0 0 0 0 0 SG fishery 6.46 9.22 1.07 1.88 2.88 3.79 3.09 6.64 7.08 3.10 1.74 1.99 3.64 4.68 2.08 1.00 1.00 5 5 6 6 5 7 7 − − − − − − − 10 10 10 10 10 10 10

× × × × × × ×

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 Demersal shark 2.2 4.4 8.4 1.4 4.0 1.0 7.6 6 5 7 6 5 5 4 5 9 6 7 − − − − − − − − − − − 10 10 10 10 10 10 10 10 10 10 10

× × × × × × × × × × ×

fishery.

0 0 0 0 0 0 0 0 0 0 0 0 GAB trawl 8.79 2.09 3.70 2.23 2.22 3.73 6.39 5.20 4.10 1.13 1.37 sardine

5 5 5 4 6 6 − − − − − − 10 10 10 10 10 10

× × × × × ×

Australian

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 SE trawl 1.52 1.40 7.84 5.83 6.04 1.59 South

=

bait pole

SASF

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 SBT and

model.

seine 5 5

− − 10 10

Ecopath

× ×

purse

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 SBT 3.00 3.00 EGAB

6 7 5 5 3 5 3 4 3 6 balanced

− − − − − − − − − − 10 10 10 10 10 10 10 10 10 10

the

× × × × × × × × × ×

in

0 0 0 0 0 0 0 0 0 0 0 0 0 SAMS-net fisheries 4.63 2.33 5.28 4.84 1.11 1.14 1.01 1.61 1.24 2.96 used

group

5 4 3 5 3 6 4 − − − − − − − 10 10 10 10 10 10 10

× × × × × × ×

functional

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 SAMS-line fisheries 5.33 1.35 1.34 8.20 2.05 1.23 1.75 by

) 2 −

km

t

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 SASF (catch

piscs invert discards

ruffs

fleet

sharks

skates

and demersal demersal

lion feeders

of grazer small

sharks

squid name

tunas-kingfish

demersal demersal demersal bentho-pelagic feeders

and sea mackerel

mackerel 0

3

omnivore planktivores (megabenthos) feeders piscs invert piscs Group Aust Pelagic Demersal Rays SBT Other Large Blue Jack AnchovySardineInshore Salmons Medium 0 Small Medium 0 Small Small Arrow Calamary Benthic Filter Sum 0 Table Summary

46 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57

monthly upwelling anomalies, calculated from the alongshore increase in catch, and conversely for increasing trophic level (Pauly

component of wind stress at South Neptune Island (Middleton and Palomares, 2005). In general, the index increases if the underly-

et al., 2007), were considered as potential PP forcing functions ing fishery expands beyond its traditional fishing area or ecosystem,

(FF) in the EGAB ecosystem. Wind stress values between 1991 and and decreases if the geographic area contracts, or if the underlying

2008 were scaled to positive values, and then converted to values food web is collapsing (Pauly and Palomares, 2005).

relative to that of January 1991. Using the anomaly search function

within Ecosim, we set the PP variance to 50 to enable the model to 2.5. Model scenarios

capture changes in the magnitude of ‘spikes’. The number of spline

We used the fitted Ecosim EGAB model to explore the

points was set to 18, to track annual anomalies in the time series.

species/group responses (increase/decrease in biomass) to reduc-

Model-generated time series were compared to the upwelling

tions in sardine biomass (to 0.75, 0.50 and 0.25 of 2008 biomass).

index time series for agreement.

The fishery effort time series was extended to year 2040, main-

taining the fishing effort of all fleets except sardine beyond 2008

2.4. Ecosystem indicators at 2008 levels, and increasing the sardine fishery effort to drive

reductions in sardine biomass. The time series of sardine fishing

After the model fitting procedure in Ecosim, we examined four mortality was deselected in these procedures. In these scenarios,

ecosystem indictors that can be used to evaluate changes in the both NZ and Australian fur seal biomass was forced to increase

marine ecosystem. (1) total catch; (2) Kempton’s index of biodiver- between 2008 and 2040 (in line with observed increases between

sity (Q), which expresses biomass species diversity of functional 1991 and 2008), so that population abundances in the EGAB were

groups with a trophic level (TL) of three or higher (Ainsworth and about 1.5 and 7.0 times present levels (2008), respectively. This is

Pitcher, 2006; Kempton and Taylor, 1976); (3) the mean trophic equivalent to an increase from ∼73,000 to ∼111,000 for NZ fur seals

∼ ∼

level of the catch (mTLC) which is calculated as the weighted aver- and 2800 to 20,000 for Australian fur seals between 2008 and

age of the TL of fishery targeted species (Pauly et al., 1998); and (4) 2040. In addition, simulations with and without fishing effort on

the Fishing in Balance Index (FIB index), which assesses whether SBT were performed by adjusting effort in the purse seine fishery

catch rates are in balance with ecosystem trophic production due beyond 2008. Scenarios where arrow squid biomass was reduced

to catch at a given TL being related to the assimilation efficiency by 50% and 75% were also examined. For all scenarios, we compared

of the ecosystem (Coll et al., 2009). The FIB index will remain con- the relative change in biomass between 2008 and 2040 with and

stant if a decline in mTLC is matched by an ecologically appropriate without PP forcing.

Table 4

Biological parameters by functional group of the balanced EGAB Ecopath model. Parameters in bold were estimated by the model. P/B = production/biomass;

Q/B = consumption/biomass; EE = ecotrophic efficiency.

−2 −1 −1

Group name Trophic level Biomass (t km ) P/B (year ) Q/B (year ) EE

1 Baleen whales 3.01 0.0389 0.020 5.097 0.000

2 Bottlenose dolphin 4.61 0.00611 0.080 16.566 0.000

3 Common dolphin 4.66 0.0039 0.090 20.511 0.000

4 NZ fur seal 4.80 0.00453 1.184 47.526 0.944

5 Aust fur seal 4.53 0.00047 1.157 28.819 0.983

6 Aust sea lion 4.91 0.00422 0.792 29.445 0.150

7 Little penguin 4.71 0.000698 1.250 85.600 0.994

8 Petrels 4.09 0.00306 1.000 147.100 0.070

9 Gannets 5.01 0.0000308 1.000 138.300 0.000

10 Terns 4.52 0.00000635 1.000 89.900 0.000

11 Pelagic sharks 4.92 0.0459 0.200 1.200 0.900

12 Demersal sharks 3.92 0.307 0.180 1.800 0.326

13 Rays and skates 3.68 0.459 0.350 2.700 0.014

14 SBT 4.52 0.145 0.200 1.600 0.900

15 Other tunas-kingfish 4.50 0.0769 0.200 1.200 0.900

16 Large bentho-pelagic piscs 4.68 0.452 0.338 3.315 0.959

17 Blue mackerel 3.23 0.219 0.370 3.500 0.857

18 Jack mackerel 3.22 0.919 0.470 3.300 0.900

19 Redbait 3.38 0.561 0.740 2.800 0.900

20 Anchovy 3.63 1.272 0.700 5.040 0.555

21 Sardine 3.36 1.517 1.600 5.040 0.330

22 Inshore small planktivores 3.93 0.489 1.010 7.300 0.900

23 Salmons and ruffs 4.51 0.246 0.440 5.400 0.900

24 Medium demersal piscs 3.47 0.302 0.485 5.400 0.844

25 Small demersal piscs 2.66 1.467 0.853 5.367 0.412

26 Medium demersal invert feeders 4.00 0.0786 0.860 5.400 0.960

27 Small demersal invert feeders 3.53 0.149 1.090 5.500 0.900

28 Mesopelagics 3.07 0.106 1.005 6.673 0.900

29 Small demersal omnivore 3.77 0.170 0.840 16.000 0.957

30 Arrow squid 4.18 0.341 1.950 3.900 0.900

31 Calamary 4.50 0.0837 1.950 3.900 0.900

32 Other squids 3.14 0.111 2.500 5.850 0.900

33 Octopus 4.16 0.294 2.500 5.850 0.900

34 Large zooplankton (carnivores) 2.20 1.287 20.000 70.000 0.800

35 Small zooplankton (herbivores) 2.00 35.119 5.000 32.000 0.800

36 Benthic grazer (megabenthos) 3.24 11.013 1.600 6.000 0.800

37 Detritivore (infauna - macrobenthos) 2.52 30.983 1.600 6.000 0.800

38 Filter feeders 2.80 1.581 1.600 6.000 0.800

39 Primary production 1.00 14.900 745.000 0.000 0.108

40 Detritus 1.00 10.0 0.009

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 47

3. Results feeders, pelagic sharks and sea lions negatively; and sea lions

affected rays and skates and little penguins negatively; NZ fur

3.1. Trophic structure and flow seals affected little penguins and petrels negatively (Fig. S1). Of the

fisheries, the SA marine scalefish net fishery affected pelagic and

The balanced model parameters of the 40 functional groups demersal sharks negatively; the SBT purse seine fishery affected

within the Ecopath model are presented in Table 4. The balanc- SBT, other tuna and kingfish negatively; the demersal shark fishery

ing procedure required adjustment to the diets of some groups affected sea lions and demersal sharks negatively, and the Spencer

where ecotrophic efficiencies (EE) were initially >1. EE is the pro- Gulf prawn fishery affected rays and skates negatively (Fig. S1).

portion of production that is either harvested or predated upon Based on this analysis, the sardine fishery has negligible impacts

by higher trophic levels and cannot exceed 1. The main dietary on other groups. However, as a steady-state analysis, all these

adjustments included reduction in the contribution of little pen- assessments do not take into account the changing abundances or

guins and petrels in the diets of NZ fur seals, and of petrels in the diets of groups. Such dynamics are explored in Ecosim scenarios

diets of Australian sea lions; reduction in the contribution of large below.

bentho-pelagic fish in the diet of arrow squid; reduction in the con- The proportional breakdown of the consumption by predators of

tribution of medium demersal invertebrate feeders in the diets of sardines and of all combined small pelagic fish groups is presented

demersal sharks and octopus; and reductions in the contribution in Fig. 4. The key predators of sardines in ranked order were large

of small demersal omnivores in the diets of salmon and ruffs and of bentho-pelagic fish, arrow squid, salmons and ruffs, SBT, other tuna

medium demersal piscivorous fish. The trophic flows between the and kingfish, common dolphins and pelagic sharks. For all small

functional groups in the EGAB ecosystem estimated by Ecopath, are pelagic fish, the key predators in ranked order were salmons and

summarised in Fig. 3. ruffs, arrow squid, SBT, calamary, other tuna and kingfish, common

Results from the mixed trophic impacts routine indicate the dolphins, little penguins, NZ fur seals, petrels and pelagic sharks

positive effect of primary production on small and large zoo- (Fig. 4). The total consumption of sardines estimated by the Ecosim

−1

plankton, the positive effects of these groups on small pelagic model was 123,369 t y , representing about a third (31%) of the

−1

fishes which in turn have a positive effect on dolphin, seal and total combined consumption of small pelagic fish (396,129 t y ) in

seabird groups (Supplementary Information, Fig. S1). Groups con- the EGAB ecosystem.

taining commercially targeted species influenced their respective

fishing fleets positively, and most groups affected themselves 3.2. Time-series fitting

negatively (Fig. S1). Cephalopod groups affected small demersal

invertebrate feeders and meso-pelagic fish negatively; salmons Ecosim model fits to fishery time series and model-derived

and ruffs affected medium and small demersal piscivorous fish wind-stress anomaly forcing functions (Fig. S2) are presented in

and inshore small planktivores negatively; and bentho-pelagic Fig. 5 There was a high degree of variability in the degree to which

piscivorous fish affected small pelagic fish and octopus groups Ecosim reproduced the CPUE (biomass) and yield (catch) trends

negatively (Fig. S1). SBT affected other tunas and kingfish nega- of various functional groups (Fig. 5). Trends in modelled biomass

tively; demersal sharks affected medium demersal invertebrate fitted observed trends reasonably well for pelagic sharks, demersal

Fig. 3. Food web diagram of the Ecopath model of the EGAB ecosystem. Functional groups are indicated by circles, with size proportional to their biomass. The trophic links

between groups are indicated by lines. Ovals (dashed) encircle major trophic groups including marine mammals, seabirds, pelagic sharks, large pelagic fish and cephalopods.

Small Pelagic fish group are encircled by a solid oval.

48 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57

Fig. 4. Estimated proportional breakdown of total annual consumption of sardines and of all small pelagic fish (sardine, anchovy, blue mackerel, jack mackerel, redbait and

inshore planktivores) by their predators based on the Ecopath model of the EGAB ecosystem. Total consumption of sardines and of all small pelagic fishes is estimated to be

−1 −1

123,369 t y and 396,129 t y , respectively.

sharks, large bentho-pelagic piscivorous fish, sardine, medium functional groups composed of species targeted by fisheries, such

demersal piscivorous fish, and calamary (Fig. 5). Trends in modelled as demersal sharks (gummy and school shark), SBT, large bentho-

catch fitted observed trends reasonably well for pelagic sharks, pelagic fish (snapper), sardine, medium demersal piscivorous fish

demersal sharks, rays and skates, SBT, other tuna and kingfish, large (King George whiting), calamary and benthic grazers (western king

bentho-pelagic piscivorous fish, jack mackerel, sardine, medium prawn); and poorer for non-targeted (byproduct) species such as

demersal piscivorous fish, small demersal piscivorous fish, arrow blue and jack mackerel and small demersal piscivorous fish (Fig. 5).

squid and benthic grazers (Fig. 5). In many instances the fit of mod- A summary of the estimated changes in the biomass of the

elled trends to observed trends was substantially better for either EGAB ecosystem between 1991 and 2008 summarised into nine

biomass or catch. For example for SBT, jack mackerel, sardine and basic groups, with and without PP forcing in the Ecosim model,

arrow squid, the fit to catch was much better than the fit to biomass; is presented Table 5. With no change in PP, the observed increase

whereas for pelagic sharks, medium demersal piscivorous fish and in high trophic predators (fur seals) over the period occurs to the

calamary, the fit to biomass was much better than the fit to catch detriment of large fish (SBT and other tuna and kingfish mainly

(Fig. 5). The latter instances were predominantly for functional due to over fishing) and of medium and small fish. However, the

groups caught across multiple fishing fleets, and because Ecosim CPUE/biomass trends for many functional groups (pelagic shark,

uses only one fleet’s CPUE time series to estimate biomass (but demersal shark, large bentho-pelagic fish, sardine, salmons and

uses all fleet effort data to estimate catch), such variability in fits ruffs, medium demersal piscivorous fish, arrow squid, calamary and

is not surprising. For SBT, the capacity for Ecosim to fit to changes benthic grazers, Fig. 5) suggest that changes in fishing effort alone

in biomass was confounded by the transition from pole and bait as are not enough to capture the observed biomass trends in these

the main fishing method to purse seining, with the pole and bait groups, and that the best model fits with minimum SS resulted

fishery ending in 2000. CPUE from the purse seine fishery was used from PP forcing, suggesting that an overall increase in PP is also

to drive biomass trends for this group. The congruence between required to support the observed biomass increases in many of the

modelled and observed trends in biomass and catch was best for groups between 1991 and 2008 (P < 0.0001, Table 5 and Fig. 5).

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 49

Fig. 5. Time series fits of the EGAB Ecosim model (thin line) to observed biomass (CPUE) and catch data (dots) for 16 functional groups between 1991 and 2008. The modelled

2

trend lines (dashed) are provided, together with values of the slope and coefficient of variation (r ) (see overleaf for remaining plots).

3.3. Temporal changes in group biomass fur seals, Australian sea lions, and bottlenose dolphins (linear

regression, P < 0.0001) but there were no significant changes in

The sardine fishery grew significantly between 1991 and 2008, the biomass of little penguins and petrels (Fig. 6A). Pelagic and

with the major growth period occurring since 2000 (Fig. 2). Land- demersal shark biomass increased significantly (linear regression,

ings of SBT increased by 75% between 1991 and 2008 (linear P < 0.0001) when fishing mortality was reduced as a result of reduc-

regression, P = 0.007), while landings from the marine scalefish net tions in effort in the marine scalefish net fishery and the demersal

and demersal shark fisheries declined by 88% and 35%, respec- shark fishery (Fig. 6B). Rays and skates also increased but pre-

tively (linear regression, P < 0.001 and P = 0.0015, respectively). dicted trends are at odds with CPUE data, especially between 2002

There has been little change in the landings from the marine and 2008 (Fig. 5). Biomass of large bentho-pelagic piscivorous

scalefish line fisheries (+7%, linear regression, P = 0.1750) and west- fish was also predicted to increase (linear regression, P < 0.0001),

ern king prawn (−1%, linear regression, P = 0.741, three regions while the biomass of SBT and other tuna and kingfish groups

combined); although GAB trawl fishery landings declined by 35%, declined (linear regression, P < 0.0001), principally in response to

the decrease was not significant (linear regression, P = 0.085) high fishing mortality. All the small pelagic fish groups (blue

(Fig. 2). mackerel, jack mackerel, redbait, anchovy, sardine, inshore small

The most significant increases in population sizes of apex preda- planktivores, salmon and ruffs) and squids (arrow squid, calamary)

tors between 1991 and 2008 were for NZ fur seals, Australian were estimated to increase in biomass between 1991 and 2008

Table 5

Change of estimated relative biomass of key taxa groups based on Ecosim simulations of the EGAB ecosystem between 1991 and 2008, without and with primary production

(PP) forcing; and between 2008 and 2040 with no PP forcing, with SBT fishery catch removed and with a 50% reduction in arrow squid biomass.

Summary groups Relative biomass change of groups

1991–2008 1991–2008 2008–2040 2008–2040 2008–2040

No PP forcing PP forcing No PP forcing No PP forcing No PP forcing

No SBT fishery No SBT fishery

<50% squid biomass

Marine mammals, birds, pelagic sharks 54% 87% 62% 85% 80%

Demersal sharks and rays 4% 35% 3% 3% 4%

Large piscivorous fish −13% 38% 2% 90% 96%

Cephalopods 4% 95% −7% −7% −23%

Medium-small fish −4% 78% 1% 1% 1%

Small and meso-pelagic fish 25% 25% −27% −27% −3%

Benthic grazers/filter feeders/detritovores 0% 30% 1% 1% 1%

Zooplankton 0% 37% 1% 1% 1%

Primary production 0% 16% 0% 0% 0%

50 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57

Fig. 6. Estimated changes in the biomass of Ecosim EGAB model groups between Fig. 7. Ecosystem indicators calculated from the Ecosim EGAB model for the period

1991 and 2008: (A) marine mammals and birds (biomass change in NZ and Aus- 1991 to 2008. (A) Kempton’s Q biomass diversity index; (B) Mean trophic level of the

tralian fur seals is forced in the model as these estimates are based on empirical catch and (C) Fishing in balance (FIB) index. Estimated trends are given by dashed

2

data for these species in the EGAB ecosystem); (B) sharks, rays and skates, tunas and lines (regression), with their coefficient of variation (r ) and level of significance (P).

kingfish; and (C) small pelagic fish, arrow squid and calamary.

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 51

(linear regression, P < 0.0001) (Fig. 6C). Common dolphins were 2008 (Fig. 7A), commensurate with increasing biomasses of marine

estimated to have declined between 1991 and 2004, with some mammals, sharks, and some large piscivorous fish and squid (Fig. 6).

recovery occurring since 2004 when management measures were The mTLC decreased significantly between 1991 and 2008 (linear

introduced to mitigate bycatch in sardine fishery (Hamer et al., regression, P < 0.001) (Fig. 7B), as a consequence of growth in the

2008). small pelagic fishery for sardines. The fishing in balance (FIB) index

between 1991 and 2008 generally increased but not significantly

3.4. Ecosystem indicators and all values were <0.02 (Fig. 7C). There was a temporary spike in

FIB in 2005 that corresponds to the peak in sardine catch in that

The ecosystem indicators identified significant changes in the year (40,000 t, Fig. 2A), but it then declined back to similar levels

EGAB ecosystem between 1991 and 2008 (Fig. 7A–C). Total catch to between 2002 and 2004 (Fig. 7C). An FIB index <0 occurs where

of the combined fisheries showed a 4.6-fold increase, attributed fishing impacts are so high that the ecosystem function is impaired,

to growth in the sardine fishery (Fig. 2A). The total catch of all or where discarding occurs and it is not considered in the analysis

other fisheries combined showed a slight (0.89) but non-significant (Christensen, 2000). An FIB index = 0 indicates high production at

reduction over the same period (Fig. 2A). Kempton’s Q biodiversity lower trophic levels with fishing in balance; and an FIB index >0

index usually increases with growing biomass of high trophic level indicates an expansion of fishing and/or where bottom-up effects

species, and decreases with increased fishing impacts. The index are occurring, resulting in more catch than expected (Coll et al.,

for the EGAB ecosystem increased significantly between 1991 and 2009).

Fig. 8. Predicted change in the biomass of functional groups in the Ecosim EGAB model in 2040 relative to 2008 with sardine biomass reduced to 0.75, 0.50 and 0.25 of 2008

biomass. In this simulation, NZ fur seal and Australian fur seal groups were forced to increase between 2008 and 2040 (in line with observed increases between 1991 and

2008), so that population abundances in the EGAB were about 1.5 and 7.0 times present (2008) levels, respectively (with no change in sardine biomass). Note: the break in

the scale of the x-axis.

52 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57

3.5. Model scenario results whose biomass declined by 6%, 17%, 27%, respectively (Fig. 8).

Little penguin decline was less affected by sardine reductions with

Functional groups that declined below their 2008 biomass declines of 12.0%, 12.5% and 12.9% in response to 25%, 50% and 75%

as a result of reductions in sardine biomass included: common reductions in sardine biomass, respectively. Relative biomasses of

dolphins, little penguins, gannets, terns, SBT, other tunas and king- NZ fur seals (65%, 61%, and 56%) and Australian sea lions (28%, 20%

fish, small demersal invertebrate feeders, mesopelagics, and arrow and 14%), were lower, respectively in response to 25%, 50% and 75%

squid (Fig. 8). For some declining groups, their decline was lessened reductions in sardine biomass, due mainly to indirect interactions

as fishing on sardines increased (e.g. petrels and jack mackerel) reducing the biomass of other key prey taxa (arrow squid and large

(Fig. 8). For some groups with increasing biomass, the increase was bentho-pelagic piscivorous fish, respectively) (Fig. 8).

lessened (e.g. NZ fur seals, Australian sea lions, pelagic sharks, and Our simulations of continuation of the current (2008) purse

salmon and ruffs), or enhanced (e.g. bottlenose dolphins, Australian seine fishing effort on tuna beyond 2008 resulted in a 21% decline

fur seals, redbait, anchovy, small demersal omnivores, other squids in biomass by 2040, on top of the estimated 51% decline up to 2008.

and octopus) as sardine biomass was reduced (Fig. 8). Simulations showed SBT biomass increased when fishing effort

Of the land-breeding marine predators, those most directly was reduced by 50%, 75% and 100% from 2008 levels, resulting

impacted by reductions in sardine biomass were terns, whose in increases in SBT biomass of 45%, 148% and 615%, respectively

biomass declined by 9%, 20% and 35% in response to 25%, 50% between 2008 and 2040 (Fig. 9). They also resulted in marked

and 75% reductions in sardine biomass, respectively; and gannets increases in the biomass of the other tuna and kingfish group (58%,

Fig. 9. Predicted change in the biomass of functional groups in the Ecosim EGAB model in 2040 relative to 2008 with biomass increases of SBT. Biomass increases in SBT

were simulated by reducing effort in the purse seine fishery by adjusting time series values between 2008 and 2040 under different scenarios of fishing effort (current [2008]

levels; 50% 2008 levels; 25% 2008 levels and no fishing effort). In these simulations, NZ fur seal and Australian fur seal groups were forced to increase between 2008 and

2040 (in line with observed increases between 1991 and 2008), so that populations abundances in the EGAB were about 1.5 and 7.0 times present (2008), respectively.

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 53

199% and 1159%, respectively). Recovery of SBT was coincident with The potential role of reduced competition for small pelagic fish

increases in the biomass of pelagic sharks, but does not have major by a reduction in biomass of arrow squid was explored by gradu-

positive or negative effects on any other groups (Fig. 9). Although ally increasing fishing mortality in the time series between 2008

the consequences are not examined here, the global quota of SBT and 2040, while excluding SBT fishing effort and enabling fur seals

was coincidentally cut by 20 per cent for the 2010 and 2011 fishing to recover as above. Results from scenarios where arrow squid

seasons since completion of this study. biomass was reduced by 50% and 75% are presented in Fig. 10, and

Simulations with and without fishing effort on SBT, and with and relative changes in the biomass of the EGAB ecosystem summari-

without PP forcing indicated that the recovery of apex predators sed into nine basic groups between 2008 and 2040 are presented

(marine mammals, birds and pelagic sharks) and of large pelagic in Table 5. Results indicate that for some high trophic-level preda-

fish such as SBT is supported through consumptions of low and tor groups that increased in biomass over the 2008–2040 period

mid-tropic species, especially small pelagic fish and cephalopods and were important predators of arrow squid (NZ fur seals, pelagic

(Table 5). The main small pelagic fish group impacted is jack mack- sharks, SBT, and other tunas and kingfish), rates of increase were

erel, with negligible change in other small pelagic species, while reduced commensurate with reductions in arrow squid biomass.

the most negatively impacted cephalopods include arrow squid and However, most others, including common dolphins, Australian

calamary (Fig. 10). fur seals, Australian sea lions, little penguins, petrels, gannets,

Fig. 10. Predicted change in the biomass of functional groups in the Ecosim EGAB model in 2040 relative to 2008 with different scenarios of arrow squid biomass reduction

and with no SBT fishery. Biomass reduction in arrow squid were simulated by adjusting time series values between 2008 and 2040 under different scenarios of fishing effort

(current [2008] levels; 50% biomass reduction of 2008 levels; 75% biomass reduction of 2008 levels). In these simulations, NZ fur seal and Australian fur seal groups were

forced to increase between 2008 and 2040 (in line with observed increases between 1991 and 2008), so that populations abundances in the EGAB were about 1.5 and 7.0

times present (2008), respectively.

54 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57

terns, and demersal sharks, responded positively to reductions in when catches were very low, and does not take into account the

arrow squid biomass (Fig. 10). Small pelagic fish species (with the substantial increases in catch and the changing abundances or

exception of inshore small planktivores), large bentho-pelagic pis- diets of groups. In contrast, dynamic assessment using Ecosim

civorous fish and salmon and ruffs also responded positively to indicated that many groups were sensitive to changes in sardine

reductions in arrow squid biomass (Fig. 10). Comparison between biomass.

the scenarios of increasing fur seal biomass with removal of SBT Of the land breeding marine predators, crested tern popula-

fishing effort, and the same scenario where arrow squid biomass tions demonstrated the greatest sensitivity to reduction in sardine

was reduced by 50% (last two columns of Table 5) highlight the biomass, both in direction (negative) and magnitude, followed by

significant predation pressure that cephalopods (especially arrow Australasian gannets. The latter species only breeds at one site off

squid) probably place on small pelagic fish, and how with reduced Cape Jaffa some distance from the centre of the sardine fishery. In

arrow squid biomass, greater production of small pelagic fish can contrast, there are at least 10 breeding colonies of crested terns

be directed into higher trophic levels (in this instance, the increased situated adjacent to the sardine fishery in southern Spencer Gulf

production of small pelagic fish was directed into large piscivorous and Investigator Strait (McLeay et al., 2010). Demographic stud-

fish and other apex predator species (Table 5)). ies of crested terns found that they were physically smaller and

had lower survival rates in years following the two sardine mass

mortality events, in 1995 and 1998 (McLeay et al., 2009b). Their

4. Discussion dependency on small pelagic fish, including sardines which they

take from waters near colonies during a short breeding season

This study presents the first trophic model of the EGAB ecosys- (McLeay et al., 2009a, 2010), gives this species the greatest potential

tem. Ecopath with Ecosim was used to resolve the trophodynamics to provide ecological performance indices for the sardine fishery,

of the EGAB ecosystem, to improve our understanding of the through foraging, reproductive and population response variables;

trophic interactions of the major functional groups and to assess the and further research to assess their suitability is warranted.

impacts of temporal changes in fishing activity and apex predator Little penguins also demonstrated a slight reduction in biomass

biomass between 1991 and 2008. The model results suggest that in response to reduced sardine biomass, although the magnitude

most apex predator populations have increased over this period, of this response was small. In Victoria, the reproductive success of

including baleen whales, bottlenose dolphins, fur seals, pelagic little penguins decreased coincident with a reduction in the con-

sharks, large bentho-pelagic piscivorous fish, small pelagic fish tribution of sardines in their diet due to a sardine mortality event

and cephalopods. Fur seal recovery is likely to have been driven (Chiaradia et al., 2003; Dann et al., 2000). The Ecosim model identi-

by a demographic response to reduced mortality from the 1970s fied that the biomass increase of New Zealand fur seals reduced in

onwards that enabled remnant populations to overcome popula- response to a reduction in sardine biomass. The Ecosim model also

tion reductions from colonial sealing that occurred 140–170 years identified common dolphins, pelagic sharks, SBT, and other tuna

earlier (Ling, 1999). In contrast, the recovery of pelagic and demer- and kingfish as sensitive to biomass reductions in sardines, but as

sal sharks, large bentho-pelagic piscivorous fish and some other some are subject to their own fishery related mortality, are less

(non-tuna) fishes indicated by the model, appears primarily to be suitable as indicator taxa compared to the land breeding marine

driven by reductions in fishing mortality. This may partially be predators.

related to reductions in domestic tuna fishing TACs since poling Ecosystem indicators identified significant changes in the EGAB

ended, and the expulsion of the Japanese/international longlining ecosystem between 1991 and 2008, the most significant of which

tuna fleets from within Australia’s EEZ including the GAB where was the 4.6-fold increase in total catch, which was entirely

they fished through the late 80s and early 90s. However, the Ecosim attributable to growth in the sardine fishery within southern

models suggest that positive PP forcing over the period improves Spencer Gulf and the Investigator Strait (Figs. 2A and 7A–C). Not

model fits, and therefore may have underpinned uniform increases surprisingly, this led to a significant reduction in the mTLC between

in small pelagic fish and cephalopod biomass that have contributed 1991 and 2008. Usually such reductions demonstrate increased

to the broad scale increase in these higher trophic levels (Fig. 9). A ecosystem impact as a consequence of a reduction in high trophic

study by Brown et al. (2010) simulated positive increases in PP to level relative to lower trophic level organisms (Coll et al., 2009).

examine the ecosystem responses to climate change scenarios in a However, the significant increase in Kempton’s Q biodiversity index

range of marine ecosystems in Australia. As with the EGAB model, over the period indicates increased growth in high trophic level

they found that such increases in PP also generally led to a posi- species, suggesting that reduction in the mTLC is attributable to

tive increase to most functional groups and to increased biomass growth in the sardine fishery, and not a reduction in high trophic

of high trophic species (Brown et al., 2010). level biomass. Furthermore, FIB index values generally close to zero

In contrast, the projected decline in SBT and other tunas and suggest high production at low trophic levels with fishing in bal-

kingfish biomass in the EGAB ecosystem appear to be driven by ance. The increasing trend in FIB index over the study period, may

increases in fishing effort. Southern bluefin tuna are a highly migra- reflect geographic expansion of fishing effort and/or more catch

tory species and the purse seine fishery in the EGAB targets juvenile than expected due to bottom-up effects such as increased primary

fish in the 3–5-year-old age classes (Wilson et al., 2009). The extent productivity (Coll et al., 2009). Values above zero indicate fishing

to which modelled changes in the EGAB reflect changes in the in balance, whereas values below zero suggest high fishing impacts

broader population, or are compounded by impacts in other juris- that are impairing ecosystem function (Christensen, 2000). The FIB

dictions and fisheries (High Seas long-line fisheries) is unclear, but index was developed in part to address the situation where a reduc-

should be investigated further. For example, the pelagic sharks and tion in the mTLC occurs as a consequence of deliberate targeting of

petrels cross several management jurisdictions within time-scales lower trophic level species, rather than due to fishing down marine

of <1 y. food webs. Another potential impact when a low trophic level fish-

The growth of the sardine fishery since its establishment in ery expands rapidly is the replacement of predation mortality with

1991 has been rapid, and its catch now exceeds that of all other fishing mortality, while overall total mortality remains relative con-

fisheries in the region by a factor of three. Despite this, sensitiv- stant. Ecosim results for the sardine fishery suggest this has not

ity analyses based on mixed trophic impacts detected negligible occurred in the EGAB ecosystem. Instead predation mortality has

impacts of the fishery on other predators. However, this steady- increased as fishing mortality and total mortality has increased

state analysis was based on the establishment years of the fishery (Supplementary Information, Fig. S3), with predation mortality as a

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 55

proportion of total mortality increasing significantly between 1991 control (Cury et al., 2000), by restricting energy flow to competi-

2

and 2008 (linear regression, P < 0.001, r = 0.616). tors at similar or higher trophic levels, although typical wasp-waist

In combination, the above indices are useful because they inte- species are highly mobile low-trophic level species, such as small

grate many aspects of fishing impacts and trophodynamic change plankton feeding pelagic fish (Bulman et al., 2010; Cury et al., 2000;

to provide measures of ecosystem health. For the EGAB ecosystem, Freon et al., 2009).

all these indices suggest that despite the large scale expansion of the There are growing concerns over the ecosystem impacts that are

sardine fishery between 1991 and 2008 and reduction in the mTLC, resulting from increased human dependence on fisheries targeting

most functional groups from cephalopods to small pelagic fish and low-trophic level species (Ainley and Blight, 2009; Branch et al.,

high trophic level predators have been increasing in biomass over 2010; Cury et al., 2011; Smith et al., 2011). Most research on the

this period. This suggests that the current fishery management impacts of these fisheries has focused on high trophic level species,

strategy is sufficiently conservative to ensure that the fishery is which are increasingly being seen to provide robust ecological per-

being managed according to the principles of ecological sustain- formance indicators of ecosystem system health, as they are often

able development (ESD) (Fletcher et al., 2002). However, the extent sensitive to subtle changes in the abundance and distribution of

to which this positive assessment reflects management of the sar- the lower trophic level species on which they prey (see Boyd et al.,

dine fishery is uncertain, given that much of the positive changes 2006 for review). Other studies have shown how these fisheries

estimated by the model reflect other changes in the ecosystem, also impact on other parts of the marine ecosystem, including other

including reductions in fishing effort and mortality in some fleets commercially targeted species (Smith et al., 2011). The growth of

(SA marine scalefish net and Commonwealth demersal gillnet shark the sardine fishery since 1991 to become Australia’s largest fishery

fishery), and a positive trend in PP over the study period. How has raised concerns over its potential ecological impacts on other

the current or alternate management strategies of the sardine fish- parts of the marine ecosystem, and the need to develop an EBFM

ery would perform under alternate oceanographic conditions (such approach. The current harvest strategy in the fishery is to maintain

as more variable upwelling), and/or in response to management a baseline total allowable commercial catch (TACC) of 30,000 t,

changes in other fisheries in the region, may be useful to explore. while estimates of the spawning biomass fall between 150,000 and

One of the key dynamics in the EGAB ecosystem is the vari- 300,000 t, corresponding to an exploitation rate of between 20%

able trends in biomass of some of the high trophic level species. and 10%, respectively (Ward et al., 2009).

Most notable is the recovery of fur seal populations over the last A recent study examined the impacts of low trophic level

25 years to numbers not seen for more than 150 years. This recov- (forage fish) depletion on global seabird populations, and identi-

ery is continuing and it is unclear at what level populations will fied a threshold prey abundance below which breeding success

stabilise. Fishing pressure on populations of SBT in the GAB and consistently declined (Cury et al., 2011). This threshold was approx-

Southern Ocean south of Australia over the last 30 years has seen imately one-third of the maximum observed prey abundance, with

these stocks reduced to a fraction of their pre-exploited levels, and the authors suggesting that a general rule of “one-third for the

they are subject to continued over fishing (Wilson et al., 2009). birds” could be applied as a guiding principal to EBFM of low-

Other important pelagic predators, such as pelagic sharks have also trophic level fisheries, by ensuring that exploitation rates maintain

been targeted by some fisheries and are subject to bycatch, and have target species above one-third of their maximum observed long-

seen historic declines. How such depletions in populations of key term biomass (Cury et al., 2011). Similarly, Smith et al. (2011)

high-trophic level predators over the last 200 years (a ‘predator- showed that broader ecosystem impacts from fishing low trophic

gap’) has affected the EGAB ecosystem is unclear, but this study level species could be substantially reduced by halving exploita-

has provided some interesting insights, in particular hypotheses tion rates from typical (∼60%) maximum sustainable yield levels to

about the dynamics between cephalopods and their predators and ∼30% exploitation rates. Based on the findings of both these stud-

competitors, and the importance of these interactions in regulating ies, the current exploitation rates in the sardine fishery (10–20%

the flow of small pelagic fish production to high trophic levels. of estimated spawning biomass and ∼24% of the total sardine con-

Ecosim scenarios that explored the ecosystem response to sumption biomass estimated by the EGAB ecosystem model) would

recovering fur seal and SBT populations out to 2040, indicated that appear conservative, and results from trophodynamic modelling

such recoveries were supported through reductions in the biomass would suggest they presently do not significantly impact ecosys-

of small pelagic fish (mainly jack mackerel) and cephalopods (espe- tem function, or high trophic level species. However, it is important

cially arrow squid). Scenarios investigating the response to reduced to highlight that most of sardine fishery catch comes from a rela-

2

arrow squid biomass as a consequence of predation and com- tively small (∼10,000 km ) area centred on southern Spencer Gulf.

petition pressure exerted by apex predators, suggest this may This area is ∼20% of the estimated spawning area of sardines in

2

enable greater production of small pelagic fish to be directed into the EGAB (∼54,000 km , Ward et al., 2009), which makes up about

higher trophic levels, in this case to large piscivorous fish and apex 7% of the total modelled area. The current estimated exploitation

predator species. This hypothesis suggests that ‘predator gaps’ that rates of sardines in the sardine fishery assumes a high degree of

resulted from reduced biomass of long-lived predator species such connectivity between the sardine population in the main fishery

as fur seal, SBT and shark, may have been quickly filled by semel- area, and other parts of the estimated spawning area within the

parous species, such as cephalopods (squids) which can build up EGAB. There currently is no data on the movements of sardines

biomass quickly in response to increased small pelagic fish avail- from the broader EGAB into the fishery to substantiate this assump-

ability and reduced predation pressure. A build-up in cephalopods tion. As such, it is possible that exploitation rates are higher within

may reduce trophic flows to high trophic levels, with their short the region of the fishery and that localised depletion cannot be

generation times and reduced predation pressure resulting in much discounted as a meso-scale issue. The model developed and pre-

of their biomass being returned to detritus. A study of the role of sented in this study covered an extensive geographic area with

small pelagic fish in southern Australian ecosystems identified that diverse habitats and may not be appropriate to examine regional

these ecosystems were largely bottom-up forced, but that differ- impacts from fisheries, such as localised depletion. This will require

ent parts of the these food webs could exert both bottom-up and the development of spatially explicit regional ecosystem models,

top-down control (Bulman et al., 2010). Furthermore, switching preferably in conjunction with appropriate ecological performance

between these states may occur in response to fishing and climate indicators (reproductive and foraging success) from land-breeding

change pressures (Bulman et al., 2010). In this respect, it is possible marine predators (e.g. seabirds and seals that forage in proximity to

that cephalopods may be able to exert some degree of ‘wasp-waist’ the fishery), to assess and identify issues of localised depletion, and

56 S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57

ensure continuing progress is made towards EBFM of the sardine Boyd, I.L., Wanless, S., Camphuysen, C.J., 2006. Top predators in marine ecosystems:

fishery. their role in monitoring and management. In: Conservation Biology. Cambridge

University Press, Cambridge.

Results from this study of the EGAB ecosystem highlight the

Branch, T.A., Stafford, K.M., Palacios, D.M., Allison, C., Bannister, J.L., Burton, C.L.K.,

importance of small pelagic fish to the higher trophic levels, the Cabrera, E., Carlson, C.A., Vernazzani, B.G., Gill, P.C., Hucke-Gaete, R., Jenner,

K.C.S., Jenner, M.N.M., Matsuoka, K., Mikhalev, Y.A., Miyashita, T., Morrice, M.G.,

trophic changes that result from loss and recovery of apex predator

Nishiwaki, S., Sturrock, V.J., Tormosov, D., Anderson, R.C., Baker, A.N., Best, P.B.,

populations, and the potential pivotal role of changing cephalopod

Borsa, P., Brownell, R.L., Childerhouse, S., Findlay, K.P., Gerrodette, T., Ilangakoon,

(squid) biomass in regulating trophic flows. The use of trophody- A.D., Joergensen, M., Kahn, B., Ljungblad, D.K., Maughan, B., McCauley, R.D.,

McKay, S., Norris, T.F., Whale, O., Rankin, S., Samaran, F., Thiele, D., Van Waere-

namic models as a tool to provide context to the potential impacts

beek, K., Warneke, R.M., Dolphin Res, G., 2007. Past and present distribution,

and management strategies of a single fishery relative to the tem-

densities and movements of blue whales Balaenoptera musculus in the Southern

poral changes of a complex ecosystem subject to dynamic impacts Hemisphere and northern . Mammal Review 37, 116–175.

from multiple fishing fleets, recovery of apex predator populations Branch, T.A., Watson, R., Fulton, E.A., Jennings, S., MacGilliard, C.R., Pablico, G.T.,

Ricars, D., Tracey, S.R., 2010. The trophic fingerprint of marine fisheries. Nature

and climate change is highlighted by this study. Although there are

468, 431–435.

clearly some limitations of the model due to data deficiencies, and

Brown, C.J., Fulton, E.A., Hobday, A.J., Matear, R.J., Possingham, H.P., Bulman, C., Chris-

the number and composition of functional groups, the model pro- tensen, V., Forrest, R.E., Gehrke, P.C., Gribble, N.A., Griffiths, S.P., Lozano-Montes,

H., Martin, J.M., Metcalf, S., Okey, T.A., Watson, R., Richardson, A.J., 2010. Effects of

vides a basis from which future improvements in model design and

climate-driven primary production change on marine food webs: implications

data inputs will enable more complex management and ecological

for fisheries and conservation. Global Change Biology 16, 1194–1212.

questions to be examined. The ecosystem performance indicators Bulman, C.M., Condie, S.A., Neira, F.J., Goldsworthy, S.D., Fulton, E.A., 2010. The

Trophodynamics of Small Pelagic Fishes in the Southern Australian Ecosystems

produced by the Ecosim model also provide a means to assess the

and the Implications for Ecosystem Modelling of Southern Temperate Fisheries.

potential impacts of the sardine fishery relative to those from other

Final Report to the Fisheries Research and Development Corporation (FRDC

fisheries and environmental change. This ability to resolve and 2008/023). CSIRO Marine and Atmospheric Research, Hobart, Tasmania.

Chiaradia, A., Costalunga, A., Kerry, K., 2003. The diet of little penguins (Eudyp-

attribute potential impacts from multiple fishing fleets and envi-

tula minor) at Phillip Island, Victoria, in the absence of a major prey – Pilchard

ronmental changes is also critical for the development and utility of

(Sardinops sagax). Emu 103, 43–48.

ecological performance indicators for assessing EBFM targets, and Christensen, V., 2000. Indicators for marine ecosystems affected by fisheries. Marine

and Freshwater Research 51, 447–450.

will not be possible without further modelling.

Christensen, V., Pauly, D., 1992. Ecopath II –a software for balancing steady-state

ecosystem models and calculating network characteristics. Ecological Modelling

Acknowledgements 61, 169–185.

Christensen, V., Walters, C.J., 2004. Ecopath with Ecosim: methods, capabilities and

limitations. Ecological Modelling 172, 109–139.

We thank the Fisheries Research and Development Corporation Christensen, V., Walters, C.J., Pauly, D., Forrest, R., 2008. Ecopath with Ecosim version

6 User Guide, November 2008. Lenfest Ocean Futures Project 2008, p. 154.

(FRDC) for support of this project (FRDC projects 2003/072, and

Coll, M., Santojanni, A., Palomera, I., Arneri, E., 2009. Food-web changes in the Adri-

2005/031), and FRDC staff for their support with project manage-

atic Sea over the last three decades. Marine Ecology Progress Series 318, 17–37.

ment, especially Crispin Ashby and Caroline Stewardson. We thank Copley, P.B., 1996. The status of seabirds in South Australia. In: Ross, G.J.B., Weaver,

Steve Shanks, Will Zacharin, Martin Smallridge (PIRSA Fisheries), K., Greig, J.C. (Eds.), The Status of Australia’s Sea birds: Proceedings of the

National Seabird Workshop. Biodiversity Group, Environment Australia, Can-

and Christian Pike project support. We thank Topa Petit (Univer-

berra, pp. 139–180.

sity of South Australia), Jeremy Austin, Sean Connell (Adelaide

Currie, D.R., Dixson, C.D., Roberts, S.D., Hooper, G.E., Sorokin, S.J., Ward, T.M., 2009.

University), Mike Steer (SARDI Aquatic Sciences) who provided co- Fishery-independent by-catch survey to inform risk assessment of the Spencer

Gulf Prawn Trawl Fishery. Report to PIRSA Fisheries. South Australian Research

supervisory support to students who assisted in this project. We

and Development Institute (Aquatic Sciences), Adelaide. SARDI Publication No.

thank Alex Ivey and Richard Saunders (SARDI Aquatic Sciences)

F2009/000369-1. SARDI Research Report Series No. 390, p. 121.

who provided additional project support, and Charles James and Currie, D.R., Sorokin, S.J., 2010. A preliminary evaluation of the distribution and

trophodynamics of demersal fish from Spencer Gulf. Report to the South

John Middleton (SARDI Aquatic Sciences) for providing wind stress

Australian Department for the Environment and Heritage. South Australian

data. Tom Okey provided help and direction during the conceptu-

Research and Development Institute (Aquatic Sciences), Adelaide. SARDI Publi-

alisation of the ecosystem modelling components of the project. cation No. F2010/000088-1. SARDI Research Report Series No. 424, p. 179.

Cury, P., Bakun, A., Crawford, R.J.M., Jarre, A., Quinones, R.A., Shannon, L.J., Verheye,

We thank Peter Gill and Margie Morrice (Blue Whale Study Inc.)

H.M., 2000. Small pelagics in upwelling systems: patterns of interaction and

for providing diet data for blue whales and Cath Kemper and Sue

structural changes in “wasp-waist” ecosystems. ICES Journal of Marine Science

Gibbs (South Australian Museum) provided diet data for bottlenose 57, 603–618.

Cury, P.M., Boyd, I.L., Bonhommeau, S., Anker-Nilssen, T., Crawford, R.J.M., Furness,

and common dolphins. We thank Peter Shaughnessy (South Aus-

R.W., Mills, J.A., Murphy, E.J., Österblom, H., Paleczny, M., Piatt, J.F., Roux, J.-

tralian Museum) for provided data on NZ fur seal pup production.

P., Shannon, L., Sydeman, W.J., 2011. Global seabird response to forage fish

We thank the following for editorial assistance: Peter Shaughnesy, depletion—one-third for the birds. Science 334, 1703–1706.

Alex Ivey, Cameron Dixon, Shane Roberts, David Currie, Mayleen Dann, P., Norman, F.I., Cullen, J.M., Neira, F.J., Chiaradia, A., 2000. Mortality and

breeding failure of little penguins, Eudyptula minor, in Victoria, 1995–96, follow-

Loo, Adrian Linnane, Jason Tanner and Tony Fowler.

ing a widespread mortality of pilchard, Sardinops sagax. Marine and Freshwater

Research 51, 355–362.

Fletcher, W.J., Chesson, J., Sainsbury, K.J., Hundloe, T., Smith, A.D.M., Whitworth, B.,

Appendix A. Supplementary data

2002. National ESD reporting Framework for Australian Fisheries: The “How to

Guide for Wild Capture Fisheries” FRDC report 2000/145, Canberra, Australia.

Supplementary data associated with this article can be found, Fowler, A.J., Lloyd, M.T., Schmarr, D., 2009. A Preliminary Consideration of By-catch

in the Marine Scalefish Fishery of South Australia, South Australian Research and

in the online version, at http://dx.doi.org/10.1016/j.ecolmodel.

Development Institute (Aquatic Sciences), Adelaide, F2009/000097-1. SARDI

2013.01.006.

Research Report Series No. 365. p. 79.

Frederiksen, M., Wright, P.J., Harris, M.P., Mavor, R.A.M.H., Wanless, S., 2005.

Regional patterns of kittiwake Rissa tridactyla breeding success are related

References to variability in sandeel recruitment. Marine Ecology Progress Series 300,

201–211.

Freon, P., Aristegui, J., Bertrand, A., Crawford, R.J.M., Field, J.C., Gibbons, M.J., Tam,

Abrams, P.A., Ginzburg, L.R., 2000. The nature of predation: prey dependent, ratio

J., Hutchings, L., Masski, H., Mullon, C., Ramdani, M., Seret, B., Simieri, M., 2009.

dependent or neither. Trends in Ecology and Evolution 15, 337–341.

Functional group biodiversity in Eastern Boundary Upwelling Ecosystems ques-

Ainley, D.G., Blight, L.K., 2009. Ecological repercussions of historical fish extraction

tions the wasp-waist trophic structure. Progress in Oceanography 83, 97–106.

from the . Fish and Fisheries 10, 13–38.

Fulton, E.A., 2010. Approaches to end-to-end ecosystem models. Journal of Marine

Ainsworth, C.H., Pitcher, T.J., 2006. Modifying Kempton’s species diversity index for

Systems 81, 171–183.

use with ecosystem simulation models. Ecological Indicators 6, 623–630.

Goldsworthy, S.D., Bulman, C., He, X., Larcombe, J., Littnan, C., 2003. Trophic inter-

Alder, J., Campbell, B., Karpouzi, V., Kaschner, K., Pauly, D., 2008. Forage fish:

actions between marine mammals and Australian fisheries: an ecosystem

from ecosystems to markets. Annual Review of Environment and Resources 33,

153–166. approach. In: Gales, N., Hindell, M., Kirkwood, R. (Eds.), Marine Mammals and

S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 57

Humans: Fisheries, Tourism and Management. CSIRO Publishing, Melbourne, Baylis, A., Morrice, M., Gill, P., McIntosh, R., Bool, N., Ward, T., 2011. The diets of

pp. 69–99. marine predators in southern Australia: assessing the need for ecosystem-based

Goldsworthy, S.D., McKenzie, J., Shaughnessy, P.D., Macintosh, R.R., Page, B., Camp- management of the South Australian sardine fishery. In: Goldsworthy, S.D., Page,

bell, R., 2009. An Update of the Report: Understanding the Impediments to B., Rogers, P., Ward, T.M. (Eds.), Establishing Ecosystem-based Management

the Growth of Australian Sea Lion Populations. Report to the Department of for the South Australian Sardine Fishery: Developing Ecological Performance

the Environment, Water, Heritage and the Arts. SARDI Publication Number Indicators and Reference Points to Assess the Need for Ecological Allocations.

F2008/00847-1, SARDI Research Report series No. 356, p. 175. Final Report to the Fisheries Research and Development Corporation, PN

Goldsworthy, S.D., Page, B., 2007. A risk-assessment approach to evaluating the sig- 2005/031.SARDI Publication No. F2010/000863-1. SARDI Research Report Series

nificance of seal bycatch in two Australian fisheries. Biological Conservation 139, No. 529. South Australian Research and Development Institute, Adelaide, pp.

269–285. 31–60.

Goldsworthy, S.D., Page, B., Shaughnessy, P.D., Linnane, A., 2010. Mitigating Seal Pauly, D., Christensen, V., Dalsgaard, J., Froese, R., Torres, F.J., 1998. Fishing down

Interactions in the SRLF and the Gillnet Sector SESSF in South Australia. Report marine food webs. Science 279, 860–863.

to the Fisheries Research and Development Institute. South Australian Research Pauly, D., Palomares, M.L., 2005. Fishing down marine food webs: it is far more

and Development Institute (Aquatic Sciences), Adelaide. SARDI Publication No. pervasive than we thought. Bulletin of Marine Science 76, 197–211.

F2009/000613-1. SARDI Research Report Series No. 405, p. 213. Pikitch, E.K., Santora, C., Babcock, E.A.A.B., Bonfil, R., Conover, D.O., Dayton, P.,

Griffiths, S.P., Young, J.W., Lansdell, M.J., Campbell, R.A., Hamption, J., Hoyle, S.D., Doukakis, P., Fluharty, D., Heneman, B., Houde, E.D., Link, J., Livingston, P.A., Man-

Langley, A., Bromhead, D., Hinton, M.G., 2010. Ecological effects of longline fish- gel, M., McAllister, M.K., Pope, J., Sainsbury, K.J., 2004. Ecosystem-based fishery

ing and climate change on the pelagic ecosystem off eastern Australia. Reviews management. Science 305, 346–347.

in Fish Biology and Fisheries 20, 239–272. Piroddi, C., Bearzi, G., Christensen, V., 2010. Effects of local fisheries and ocean pro-

Guénette, S., Christensen, V., Pauly, D., 2008. Trophic modelling of the Peruvian ductivity on the northeastern Ionian Sea ecosystem. Ecological Modelling 221,

upwelling ecosystem: towards reconciliation of multiple datasets. Progress in 1526–1544.

Oceanography 79, 326–335. Polovina, J.J., 1984. Model of a coral reef ecosystem. The ECOPATH model and its

Hamer, D.J., Ward, T.M., McGarvey, R., 2008. Measurement, management and miti- application to French Frigate Shoals. Coral Reefs 3, 1–11.

gation of operational interactions between the South Australian Sardine Fishery Roberts, S.D., Steer, M.A., 2010. By-product assessment in the Spencer Gulf Prawn

and short-beaked common dolphins (Delphinus delphis). Biological Conservation Fishery with an emphasis on developing management options for Balmain bugs.

141, 2865–2878. South Australian Research and Development Institute (Aquatic Sciences), Ade-

Jahncke, J., Checkley, D.M., Hunt, G.L., Harding, J.S., 2004. Trends in carbon flux to laide. SARDI Publication No. F2010/000165-1. SARDI Research Report Series No.

seabirds in the Peruvian upwelling system: effects of wind and fisheries on 439, p. 48.

population regulation. Fisheries Oceanography 13, 208–223. Samhouri, J.F., Levin, P.S., Harvey, C.J., 2009. Quantitative evaluation of marine

Jennings, G., McGlashan, D.J., Furness, R.W., 2012. Responses to changes in sprat ecosystem indicator performance using food web models. Ecosystems 12,

abundance of common tern breeding numbers at 12 colonies in the Firth of 1283–1298.

Forth, east Scotland. ICES Journal of Marine Science 69, 572–577. Shannon, L.J., Neira, S., Taylor, M., 2008. Comparing internal and external drivers

Kempton, R.A., Taylor, L.R., 1976. Models and statistics for species diversity. Nature in the southern Benguela and the southern and northern Humboldt upwelling

262, 818–820. ecosystems. African Journal of Marine Science 30, 63–84.

Knight, M.A., Tsolos, A., 2010. South Australian Wild Fisheries Information and Shaughnessy, P.D., McKenzie, J., Lancaster, M.L., Goldsworthy, S.D., Dennis, T.E.,

Statistics Report 2008/09. South Australian Research and Development Insti- 2010. Australian fur seals establish haulout sites and a breeding colony in South

tute (Aquatic Sciences), Adelaide. SARDI Publication No. F2008/000804-2, SARDI Australia. Australian Journal of Zoology 58, 94–103.

Research Report Series No. 417, p. 64. Shaughnessy, P.D., McKeown, A., 2002. Trends in abundance of New Zealand fur

Ling, J.K., 1999. Exploitation of fur seals and sea lions from Australian, New Zealand seals, Arctocephalus forsteri, at the , South Australia. Wildlife

and adjacent subantarctic islands during the eighteenth, nineteenth and twen- Research 29, 363–370.

tieth centuries. Australian Zoologist 31, 323–350. Smith, A.D.M., Brown, C.J., Bulman, C.M., Fulton, E.A., Johnson, P., Kaplan, I.C., Lozano-

Lozano-Montes, H.M., Loneragan, N.R., Babcock, R.C., Jackson, K., 2011. Using trophic Montes, H., Mackinson, S., Marzloff, M., Shannon, L.J., Shin, Y.-J., Tam, J., 2011.

flows and ecosystem structure to model the effects of fishing in the Jurien Bay Impacts of fishing low-trophic level species on marine ecosystems. Science 333,

Marine Park, temperate Western Australia. Marine and Freshwater Research 62, 1147–1150.

421–431. Ulanowicz, R.E., Puccia, C.J., 1990. Mixed trophic impacts in ecosystems. Coenoses

Mann, K.H., Lazier, J.R.N., 1996. Dynamics of Marine Ecosystems. Blackwell Scientific, 5, 7–16.

Oxford, p. 394. van Ruth, P.D., Ganf, G.G., Ward, T.M., 2010a. Hot-spots of primary productivity: an

McClatchie, S., Middleton, J.F., Ward, T.M., 2006. Water mass analysis and alongshore alternative interpretation to conventional upwelling models. Estuarine, Coastal

variation in upwelling strength in the eastern Great Australian Bight. Journal of and Shelf Science 90, 142–158.

Geophysical Research 111, 1–13. van Ruth, P.D., Ganf, G.G., Ward, T.M., 2010b. The influence of mixing on primary

McLeay, L.J., Page, B., Goldsworthy, S.D., Paton, D.C., Teixeira, C., Burch, P., Ward, T.M., productivity: a unique application of classical critical depth theory. Progress in

2010. Foraging behaviour and habitat use of a short-ranging seabird, the crested Oceanography 85, 224–235.

tern. Marine Ecology Progress Series 411, 271–283. Walters, C.J., Christensen, V., Pauly, D., 1997. Structuring dynamic models of

McLeay, L.J., Page, B., Goldsworthy, S.D., Ward, T.M., Paton, D.C., 2009a. Size mat- exploited ecosystems from trophic mass-balance assessments. Reviews in Fish

ters: variation in the diet of chick and adult crested terns. Marine Biology 156, Biology and Fisheries 7, 139–172.

1765–1780. Ward, T.M., Hoedt, F., McLeay, L., Dimmlich, W.F., Kinloch, M., Jackson, G., McGarvey,

McLeay, L.J., Page, B., Goldsworthy, S.D., Ward, T.M., Paton, D.C., Waterman, M., Mur- R., Rogers, P.J., Jones, K., 2001. Effects of the 1995 and 1998 mass mortality events

ray, M.D., 2009b. Demographic and morphological responses to prey depletion in on the spawning biomass of sardine. Sardinops sagax, in South Australian waters.

a crested tern (Sterna bergii) population: can fish mortality events highlight per- ICES Journal of Marine Science 58, 865–875.

formance indicators for fisheries management? ICES Journal of Marine Science Ward, T.M., Ivey, A.R., Mcleay, L.J., Burch, P., 2009. Spawning biomass of sardine,

66, 237–247. Sardinops sagax, in waters off South Australia in 2009. South Australian Research

Merino, G., Barange, M., Mullon, C.n., 2010. Climate variability and change scenarios and Development Institute (Aquatic Sciences), Adelaide. Final Report to PIRSA

for a marine commodity: modelling small pelagic fish, fisheries and fishmeal in Fisheries, SARDI Publication No. F2007/000566-3.

a globalized market. Journal of Marine Systems 81, 196–205. Ward, T.M., Mcleay, L.J., Dimmlich, W.F., Rogers, P.J., McClatchie, S., Matthews,

Middleton, J.F., Arthur, C., vanRuth, P., Ward, T.M., McClean, J.L., Maltrud, M.E., Gill, P., R., Kampf, J., van Ruth, P.D., 2006. Pelagic ecology of a northern boundary

Levings, A., Middleton, S., 2007. El Ninoˇ effects and upwelling off South Australia. current system: effects of upwelling on the production and distribution of sar-

Journal of Physical Oceanography 37, 2458–2477. dine (Sardinops sagax), anchovy (Engraulis australis) and Southern Bluefin Tuna

Middleton, J.F., Bye, J.A.T., 2007. A review of the shelf-slope circulation along (Thunnus maccoyii) in the Great Australian Bight. Fisheries Oceanography 15,

Australia’s southern shelves: Cape Leeuwin to Portland. Progress in Oceanog- 191–207.

raphy 75, 1–41. Ward, T.M., Mcleay, L.J., Rogers, P.J., 2005. Spawning biomass of sardine (pilchard,

Middleton, J.F., Cirano, M., 2002. A boundary current along Australia’s southern Sardinops sagax) in South Australian waters. South Australian Research and

shelves: the Flinders Current. Journal of Geophysical Research 107. Development Institute (Aquatic Sciences), Adelaide. Final Report prepared for

Middleton, J.F., Platov, G., 2003. The mean summertime circulation along Australia’s PIRSA Fisheries, RD05/0019-1.

southern shelves: a numerical study. Journal of Physical Oceanography 33, Ward, T.M., Rogers, P.J., McLeay, L.J., 2007. Developing and evaluating the use of the

2270–2287. Daily Egg Production Method to estimate the spawning biomass of blue mack-

Myers, R.A., Baum, J., Sheperd, T., Powers, S.P., Peterson, C.H., 2007. Cascading erel Scomber australasicus off southern and eastern Australia. In: Ward, T.M.,

Effects of the Loss of Apex Predatory Sharks from a Coastal Ocean. Science 315, Rogers, P.J. (Eds.), Development and Evaluation of Egg-based Stock Assessment

1846–1850. Methods for Blue Mackerel Scomber australasicus in Southern Australia. Report

Naylor, R.L., Hardy, R.W., Bureau, D.P., Chiu, A., Elliott, M., Farrell, A.P., Forster, I., to the Fisheries Research and Development Corporation, Project No. 2002/061,

Gatlin, D.M., Goldburg, R.J., Hua, K., Nichols, P.D., 2009. Feeding aquaculture in pp. 311–352.

an era of finite resources. Proceedings of the National Academy of Sciences of Wilson, D., Curtotti, R., Begg, G., Phillips, K., 2009. Fishery Status Reports 2008: Status

the United States of America 106, 15103–15110. of Fish Stocks and Fisheries Managed by the Australian Government. Bureau of

Page, B., Goldsworthy, S.D., McLeay, L., Wiebkin, A., Peters, K., Einoder, L., Rogers, P., Rural Sciences and Australian Bureau of Agriculture and Resource Economics,

Braley, M., Gibbs, S., McKenzie, J., Huveneers, C., Caines, R., Daly, K., Harrison, S., Canberra.