Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological...

20
Ecological Modelling 255 (2013) 38–57 Contents lists available at SciVerse ScienceDirect Ecological Modelling jo u r n al hom ep age : www.elsevier.com/locate/ecolmodel Trophodynamics of the eastern Great Australian Bight ecosystem: Ecological change associated with the growth of Australia’s largest fishery Simon D. Goldsworthy a,, Brad Page a,b , Paul J. Rogers a,c , Cathy Bulman d , Annelise Wiebkin a,e,b , Lachlan J. McLeay a,e,b , Luke Einoder a,e , Alastair M.M. Baylis a,e,f , Michelle Braley a , Robin Caines g , Keryn Daly e , Charlie Huveneers a,c , Kristian Peters a,e , Andrew D. Lowther a , Tim M. Ward a a South Australian Research and Development Institute, Aquatic Sciences, 2 Hamra Ave, West Beach, South Australia 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 Article history: Received 17 May 2012 Received in revised form 21 January 2013 Accepted 23 January 2013 Available online 1 March 2013 Keywords: Eastern Great Australian Bight Food web model Ecopath with Ecosim Sardine Sardinops sagax Fishing impacts a b s t r a c t We used the Ecopath with Ecosim software to develop a trophic mass-balance model of the eastern Great Australian Bight ecosystem, off southern Australia. Results provide an ecosystem perspective of Australia’s largest fishery, the South Australian sardine fishery, by placing its establishment and growth in the context of other dynamic changes in the ecosystem, including: the development of other fish- 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 seals and seabirds which may be suitable ecological performance indicators of ecosystem health. Results indicate that despite the rapid growth of the sardine fishery since 1991, there has likely been a negli- gible fishery impact on other modelled groups, suggesting that current levels of fishing effort are not impacting negatively on the broader ecosystem structure and function in the eastern Great Australian Bight. Results highlight the importance of small pelagic fish to higher trophic levels, the trophic changes 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 There is growing concern about the global impacts of increased fishing effort on ecosystem structure and function. Much of this concern stems from the growth of fisheries that target low trophic level species (planktivorous small pelagic fish), which currently account for 30% of global landings (Ainley and Blight, 2009; Branch et al., 2010). This growth has been largely driven by increased demand for fish meal production, feed or fertiliser for the aquaculture and agriculture industries and only a small portion is processed for human consumption (Alder et al., 2008; Merino et al., 2010; Naylor et al., 2009). Small pelagic fishes are recognised Corresponding author. E-mail address: [email protected] (S.D. Goldsworthy). as important in marine ecosystems in the transfer of production from plankton to higher trophic level species (marine mammals, seabirds and large predatory fish), and concerns about the poten- tial impacts on high trophic level species from localised depletion by fisheries has been identified by many studies (Cury et al., 2000, 2011; Jennings et al., 2012; Smith et al., 2011). Although some studies have successfully demonstrated the relationships between overfishing of low trophic level species and the impacts on higher trophic levels, especially seabirds (Cury et al., 2011; Frederiksen et al., 2005; Jahncke et al., 2004; Myers et al., 2007), the numerical and trophodynamic relationships between predators and prey are 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 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

Transcript of Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological...

Page 1: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

Tc

SLKa

b

c

d

e

f

g

a

ARRAA

KEFESSF

1

ficlaBitie

0h

Ecological Modelling 255 (2013) 38– 57

Contents lists available at SciVerse ScienceDirect

Ecological Modelling

jo u r n al hom ep age : www.elsev ier .com/ locate /eco lmodel

rophodynamics of the eastern Great Australian Bight ecosystem: Ecologicalhange associated with the growth of Australia’s largest fishery

imon D. Goldsworthya,∗, Brad Pagea,b, Paul J. Rogersa,c, Cathy Bulmand, Annelise Wiebkina,e,b,achlan J. McLeaya,e,b, Luke Einodera,e, Alastair M.M. Baylisa,e,f, Michelle Braleya, Robin Cainesg,eryn Dalye, Charlie Huveneersa,c, Kristian Petersa,e, Andrew D. Lowthera, Tim M. Warda

South Australian Research and Development Institute, Aquatic Sciences, 2 Hamra Ave, West Beach, South Australia 5024, AustraliaDepartment of Environment, Water and Natural Resources, GPO 1047, Adelaide, South Australia 5001, AustraliaSchool of Biological Sciences, Flinders University, GPO Box 2100, Adelaide, South Australia 5001, AustraliaCSIRO Marine Research, GPO Box 1538, Hobart, Tasmania 7001, AustraliaUniversity of Adelaide, School of Earth and Environmental Sciences, North Terrace, Adelaide, South Australia 5005, AustraliaDeakin University, Princess Highway, Warrnambool, Victoria 3280, AustraliaUniversity of South Australia, Mawson Lakes, South Australia 5095, Australia

r t i c l e i n f o

rticle history:eceived 17 May 2012eceived in revised form 21 January 2013ccepted 23 January 2013vailable online 1 March 2013

eywords:astern Great Australian Bightood web modelcopath with Ecosimardine

a b s t r a c t

We used the Ecopath with Ecosim software to develop a trophic mass-balance model of the easternGreat Australian Bight ecosystem, off southern Australia. Results provide an ecosystem perspective ofAustralia’s largest fishery, the South Australian sardine fishery, by placing its establishment and growthin the context of other dynamic changes in the ecosystem, including: the development of other fish-eries; changing abundances of apex predator populations and oceanographic change. We investigatedthe potential impacts of the sardine fishery on high tropic level predators, particularly land-breedingseals and seabirds which may be suitable ecological performance indicators of ecosystem health. Resultsindicate that despite the rapid growth of the sardine fishery since 1991, there has likely been a negli-gible fishery impact on other modelled groups, suggesting that current levels of fishing effort are notimpacting negatively on the broader ecosystem structure and function in the eastern Great Australian

ardinops sagaxishing impacts

Bight. Results highlight the importance of small pelagic fish to higher trophic levels, the trophic changesthat have resulted from loss and recovery of apex predator populations, and the potential pivotal roleof cephalopod biomass in regulating ‘bottom-up’ trophic processes. The ability to resolve and attributepotential impacts from multiple fisheries, other human impacts and ecological change in this poorlyunderstood region is highlighted by the study, and will be critical to ensure future ecologically sustainabledevelopment within the region.

. Introduction

There is growing concern about the global impacts of increasedshing effort on ecosystem structure and function. Much of thisoncern stems from the growth of fisheries that target low trophicevel species (planktivorous small pelagic fish), which currentlyccount for ∼30% of global landings (Ainley and Blight, 2009;ranch et al., 2010). This growth has been largely driven by

ncreased demand for fish meal production, feed or fertiliser for

he aquaculture and agriculture industries and only a small portions processed for human consumption (Alder et al., 2008; Merinot al., 2010; Naylor et al., 2009). Small pelagic fishes are recognised

∗ Corresponding author.E-mail address: [email protected] (S.D. Goldsworthy).

304-3800/$ – see front matter. Crown Copyright © 2013 Published by Elsevier B.V. All rittp://dx.doi.org/10.1016/j.ecolmodel.2013.01.006

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

as important in marine ecosystems in the transfer of productionfrom plankton to higher trophic level species (marine mammals,seabirds and large predatory fish), and concerns about the poten-tial impacts on high trophic level species from localised depletionby fisheries has been identified by many studies (Cury et al., 2000,2011; Jennings et al., 2012; Smith et al., 2011). Although somestudies have successfully demonstrated the relationships betweenoverfishing of low trophic level species and the impacts on highertrophic levels, especially seabirds (Cury et al., 2011; Frederiksenet al., 2005; Jahncke et al., 2004; Myers et al., 2007), the numericaland trophodynamic relationships between predators and prey arepoorly understood, even in regions where there are commercially

valuable fisheries (Abrams and Ginzburg, 2000; Cury et al., 2011). Assuch, numerical modelling of ecosystem dynamics is increasinglybeing utilised to resolve these complex relationships (see review byFulton, 2010) and to achieve ecosystem based fishery management

ghts reserved.

Page 2: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

ogical

(oh2smfiicwoo

de(AeGebef∼i22ct

F

S.D. Goldsworthy et al. / Ecol

EBFM) objectives (Pikitch et al., 2004). The latter is dependentn the development of reliable indicators of ecosystem status andealth (Cury et al., 2011; Lozano-Montes et al., 2011; Pikitch et al.,004; Samhouri et al., 2009). The objectives underpin ecologicallyustainable development (ESD) frameworks that form the basis forodern management of Australian fisheries; however, there are

ew examples where they are achieved or put into practice. Theynclude: (i) avoid degradation of ecosystems, (ii) minimise risk ofrreversible change to assemblages of species and ecosystem pro-esses, (iii) obtain and maintain long-term socioeconomic benefitsithout compromising the ecosystem, and (iv) generate knowledge

f ecosystem processes sufficient to understand the consequencef human actions (Pikitch et al., 2004).

The South Australian sardine (Sardinops sagax) fishery (sar-ine fishery) is Australia’s largest fishery by mass of catch. It wasstablished in 1991 to provide feed for the southern bluefin tunaSBT, Thunnus maccoyii) mariculture industry in Port Lincoln, Southustralia (Fig. 1) (Ward et al., 2005). Operated as a purse seine fish-ry, sardine catch within this region is centred on southern Spencerulf, Investigator Strait and the western Eyre Peninsula in the east-rn Great Australian Bight (EGAB) (Ward et al., 2005). Spawningiomass of sardine was estimated to be ∼165,000 t in 1995 (Wardt al., 2009), but it declined by over 70% to ∼37,000 t in 1996ollowing an unprecedented mass mortality event, recovered to146,000 t in 1998 and then declined by over 70% again to ∼36,000 t

n early 1999 following a second mass mortality event (Ward et al.,

001). Between 1994 and 2001, fishery catches ranged between500 and 6500 t each year, but the total allowable commercialatch (TACC) was expanded and the harvest increased markedlyo 42,475 t in 2005, and then stabilising around 30,000 t from 2007.

ig. 1. Location of the eastern Great Australian Bight (EGAB) ecosystem included in the E

Modelling 255 (2013) 38– 57 39

The EGAB supports high densities of small pelagic fishes,the dominant species being sardine and anchovy (Engraulisaustralis) (Ward et al., 2006). Other important small pelagicfish species include blue mackerel (Scomber australasicus), jackmackerel (Trachurus declivis), yellowtail scad (T. novaezelandiae),redbait (Emmelichthys nitidus), maray (Etrumeus teres) and saury(Scomberesox saurus). Primary and secondary production in theEGAB appears to be underpinned by a unique northern bound-ary current system, the Flinders Current (Middleton and Bye, 2007;Middleton and Cirano, 2002; van Ruth et al., 2010a, 2010b; Wardet al., 2006), the dominant oceanographic feature of which is coastalupwelling that occurs between November to May, especially offthe Bonney Coast, Kangaroo Island and southern and western EyrePeninsula (McClatchie et al., 2006; Middleton and Bye, 2007; vanRuth et al., 2010a, 2010b; Ward et al., 2006) (Fig. 1). This coastalupwelling has oceanographic, biological and ecological similari-ties to the larger and more productive eastern boundary currentsthat underpin the Benguela, Humbolt, California and Canary Cur-rent upwelling systems (Mann and Lazier, 1996; Middleton andPlatov, 2003; Ward et al., 2006), although the latter typically havean order of magnitude higher productivity and small pelagic fishbiomass compared to the EGAB (Guénette et al., 2008).

The rich pelagic resources of the EGAB underpin arguably thegreatest density and biomass of apex predators to be found inAustralian coastal waters. These include marine mammals such aspygmy blue whales (Balaenoptera musculus brevicauda), and >80%

of Australia’s populations of New Zealand fur seals (Arctocephalusforsteri) and Australian sea lions (Neophoca cinerea) (Branch et al.,2007; Goldsworthy et al., 2003, 2009), and the Australian fur seal (A.pusillus doriferus) (Shaughnessy et al., 2010). All seal species were

copath with Ecosim model. The depth contours are 100, 200, 500, 1000 and 2000 m.

Page 3: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

4 ogical

st2a(tc(akG

i(2sSatsa

tcatmlompttfiisaafseiaaos

a(hmtpdfiieim

2

2

(

0 S.D. Goldsworthy et al. / Ecol

ubjected to early colonial sealing, with recovery of fur seal popula-ions commencing in the 1970s and 1980s (Goldsworthy and Page,007; Ling, 1999; Shaughnessy and McKeown, 2002). Other keypex predators include seabirds, such as short-tailed shearwatersPuffinus tenuirostris), little penguins (Eudyptula minor) and crestederns (Sterna bergii); pelagic sharks including bronze whalers (Car-harhinus brachyurus) and dusky shark (C. obscurus), white sharkCarcharodon carcharias) and shortfin mako (Isurus oxyrinchus);nd predatory fishes, such as southern bluefin tuna and yellowtailingfish (Seriola lalandi) (Copley, 1996; Goldsworthy et al., 2003;oldsworthy and Page, 2007; Ward et al., 2006).

The EGAB supports some of Australia’s most valuable fisheries,ncluding four main Commonwealth and five main South AustralianState) managed fisheries (Knight and Tsolos, 2010; Wilson et al.,009). The main Commonwealth fisheries that operate are theouthern bluefin tuna purse seine fishery (SBT), GAB Trawl (GABT),outh East Trawl (SET), and shark gillnet component of the Gill Hooknd Trap fishery (GHAT) (Wilson et al., 2009). The main South Aus-ralian State managed fisheries that operate in the region are theardine fishery, southern rock lobster, abalone, western king prawnnd marine scalefish (MSF) fisheries (Knight and Tsolos, 2010).

In recognition of the economic importance of the sardine fishery,he important role of small pelagic fishes in underpinning ecologi-al processes in the EGAB ecosystem, and the high socio-economicnd conservation significance of the region’s marine predators,he fishing industry and fishery managers identified the need to

anage potential ecological impacts. In response to this need, aarge-scale ecosystem programme was initiated to assess the rolef sardines in the EGAB ecosystem and to develop ecological perfor-ance indicators and reference points for the fishery. As part of this

rogramme, this study aimed to: (1) develop a food web model forhe EGAB and describe its temporal dynamics over the period sincehe sardine fishery established in 1991 using time-series data ofshing activity and environmental drivers; (2) identify the trophic

nteractions and functional groups most sensitive to variability inardine biomass; (3) determine the sensitivity of land breedingpex predators to changes in sardine biomass and fishery catch, andssess their appropriateness as ecological performance indicatorsor the sardine fishery; (4) examine temporal changes to ecosystemtructure using ecosystem indicators and assess their potential ascological performance indicators of ecosystem health; (5) exam-ne the trophodynamic changes occurring in response to changes inpex predator biomasses, including recovering fur seal populationsnd declining SBT populations and (6) examine how the recoveryf these groups is likely to impact on the EGAB ecosystem and theardine fishery.

Despite global recognition of the need to adopt sustainablepproaches to the management of fisheries and aquaculturePikitch et al., 2004), there are still few examples where EBFMas been adopted to complement the existing single-species/stockanagement paradigm. The principal aim of this study was

o provide an ecosystem perspective of the sardine fishery, bylacing its establishment and growth in the context of otherynamic changes in the ecosystem, including changes to othersheries, changes to apex predator populations, and meteorolog-

cal and oceanographic changes. Such information is essential tonhance EBFM of Australia’s largest pelagic fishery by incorporat-ng scientifically-based approaches for assessing and, if necessary,

itigating the fishery’s trophic impacts.

. Materials and methods

.1. Ecopath and mass balance approach

We used the Ecopath with Ecosim software v 6.1.0www.Ecopath.org) to develop a trophic mass-balance model

Modelling 255 (2013) 38– 57

of the EGAB ecosystem. Ecopath was developed by Polovina(1984), based on a simple steady-state trophic box model, andfurther developed by Christensen and Pauly (1992) and Walterset al. (1997). Whereas Ecopath enables description of the staticstate energy flow of an ecosystem at a particular point in time,Ecosim enables dynamic simulations based on Ecopath parametersthat allow the estimation of ecosystem response to environmentalperturbations. The Ecopath with Ecosim software has now beenused to describe a diverse range of aquatic ecosystems world-wide,and details of the ecological theory and mathematical equationsthat underpin its key functions have been extensively detailedelsewhere (e.g. Christensen and Walters, 2004; Griffiths et al.,2010; Piroddi et al., 2010; Shannon et al., 2008).

2.2. Model area and structure

We modelled the eastern Great Australian Bight (EGAB) pelagicecosystem, a region off the South Australian coast that includes con-tinental shelf waters to 200 m depth between 132◦E and 139.7◦E.The model domain includes Investigator Strait and the mid-lowerportions of Gulf St Vincent and Spencer Gulf where the sardine fish-ery is centred. The total modelled area was ∼154,084 km2 (Fig. 1).This area was selected because the distribution and abundance ofsardine and other small pelagic fish species was well known fromsmall pelagic fish ichthyoplankton surveys between 1995 and 2010(Ward et al., 2007, 2009).

As the aims of the model were to investigate the potentialimpacts of the sardine fishery on high tropic level predators,especially land-breeding seals and seabirds, these groups were dis-aggregated into single species where data on diet and populationbiomass permitted. As we were also interested in the broad ecolog-ical roles and importance of small pelagic fish in general, includingsardines, sardines, anchovies, jack mackerel, redbait, and other keymid-trophic level species such as arrow squid and calamary, werealso examined as single-species functional groups where data per-mitted.

A total of 40 functional and species groups were devel-oped based on species similarity in terms of diet, habitat,foraging behaviour, size, consumption and rates of production(Supplementary Information A and B). Intrinsic to Ecopath modeldevelopment, each trophic group operates as a single biomass,despite groups often being composed of several species. The aggre-gation of species into trophic groups will therefore impact on modeldynamics in some instances; however, by matching species fordiet, consumption, and production rates we attempted to constrainthe errors and uncertainty of aggregating. The model was para-meterised for 1991, when the sardine fishery commenced. Modelsimulations were run over the following 18 years (1991–2008).For each trophic group, four inputs were derived from the avail-able data: diet, biomass, production per unit of biomass (P/B) andconsumption per unit of biomass (Q/B). A dietary matrix was con-structed using empirical data derived within the EGAB ecosystemwhere possible (Table 1). For marine mammals, seabirds, large andsmall pelagic fishes, arrow squid and calamary, dietary data summ-arised by Page et al. (2011) was used (Table 1). For most otherfish species in the region, data from Currie and Sorokin (2010) wasused. Where possible, the biomasses (t km−2) of functional groupswere estimated either from field surveys or stock assessments. Adetailed description of the functional groups and how estimates ofbiomass, P/B and Q/B were derived is presented as SupplementaryInformation (A and B).

Fishery data on landings, discards and effort between 1991 and

2008 were obtained from State and Commonwealth logbook data.A total of 11 fisheries (fleets) operate within the EGAB ecosystem(Fig. 2, Tables 2 and 3). Six were South Australian (State) managedfisheries: the South Australian (SA) sardine, SA Marine Scalefish
Page 4: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

S.D.

Goldsw

orthy et

al. /

Ecological M

odelling 255 (2013) 38– 57

41

Table 1Functional groups and their diet matrix showing proportional prey contribution used in the balanced EGAB Ecopath model.

Group name Group 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

Baleen whales 1Bottlenose dolphin 2Common dolphin 3NZ fur seal 4Aust fur seal 5Aust sea lion 6Little penguin 7 0.0015 0.0004Petrels 8 0.0010Gannets 9Terns 10Pelagic sharks 11Demersal sharks 12 0.0025 0.0009 0.0133Rays and skates 13 0.0042 0.0064SBT 14 0.0818Other tunas-kingfish 15Large bentho-pelagic

piscs16 0.0406 0.0862 0.0712 0.0640 0.0954 0.0006 0.0153 0.3850 0.0586 0.0002 0.0777

Blue mackerel 17 0.0004 0.0032 0.0011 0.0161 0.0002 0.1100 0.0043 0.0288 0.0606 0.0447 0.0303Jack mackerel 18 0.0153 0.1137 0.0611 0.0536 0.0083 0.0509 0.0016 0.0019 0.0720 0.0356 0.0360Redbait 19 0.1186 0.4489 0.0441 0.0045 0.0141 0.1818 0.0433 0.0909Anchovy 20 0.4328 0.0055 0.6152 0.0044 0.0900 0.3072 0.0781 0.0750 0.0375Sardine 21 0.0209 0.2082 0.0109 0.0010 0.0353 0.0004 0.1600 0.2294 0.2298 0.4288 0.5267 0.2164Inshore small

planktivores22 0.0063 0.0003 0.0406 0.0002 0.0075 0.0783

Salmons and ruffs 23 0.0013 0.0170 0.0145 0.0025 0.0885 0.0200 0.0249 0.0072 0.0370Medium demersal

piscs24 0.0087 0.0630 0.3674 0.1097 0.0053 0.0375 0.0567 0.0021

Small demersal piscs 25 0.2326 0.0208 0.0375 0.0029 0.0075 0.0075 0.1612 0.0004Medium demersal

invert feeders26 0.0021 0.0022 0.0007 0.1000

Small demersal invertfeeders

27 0.0140 0.0019 0.0043 0.0027 0.0161 0.0001 0.0007 0.0003

Mesopelagics 28 0.0301 0.0012 0.0161Small demersal

omnivore29 0.0467 0.0095 0.0050 0.0048 0.0151 0.1137 0.1500 0.0527 0.0001

Arrow squid 30 0.0007 0.0124 0.2697 0.0460 0.0475 0.0273 0.0104 0.0200 0.0021 0.1904 0.1818 0.1341 0.0909Calamary 31 0.1337 0.0602 0.0022 0.1472 0.0200 0.0037 0.2166 0.1187Other squids 32 0.2004 0.0248 0.0008 0.0034 0.1905 0.0271 0.0002 0.0814 0.0232 0.0502Octopus 33 0.2846 0.0029 0.0013 0.0051 0.3029 0.0004 0.0603 0.3771Large zooplankton

(carnivores)34 0.0250 0.0098 0.0964 0.0045 0.0147 0.0076 0.0137 0.0163

Small zooplankton(herbivores)

35 0.4750 0.0361 0.0237 0.0009 0.0777 0.0108 0.7604 0.7438 0.5799

Benthic grazer(megabenthos)

36 0.0004 0.0468 0.0055 0.0082 0.0029 0.0012 0.2790 0.2219 0.0234 0.1834 0.1197 0.1960

Detritivore (infauna -macrobenthos)

37 0.0524 0.7727 0.0092 0.1164 0.0073

Filter feeders 38 0.2032 0.0118 0.0292 0.0064 0.1841Primary production 39 0.1264 0.0002 0.0194 0.0163Detritus 40Import 41 0.5000 0.2979 0.8161

Page 5: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

42S.D

. G

oldsworthy

et al.

/ Ecological

Modelling

255 (2013) 38– 57

Table 1 (Continued)

Group name Group 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38

Baleen whales 1Bottlenose dolphin 2Common dolphin 3NZ fur seal 4Aust fur seal 5Aust sea lion 6Little penguin 7Petrels 8Gannets 9Terns 10Pelagic sharks 11Demersal sharks 12Rays and skates 13SBT 14Other tunas-kingfish 15Large bentho-pelagic

piscs16 0.0400 0.0990

Blue mackerel 17 0.0016Jack mackerel 18 0.1032 0.3940Redbait 19 0.0420 0.0744Anchovy 20 0.1920 0.0371 0.1120Sardine 21 0.0980 0.1213Inshore small

planktivores22 0.3251 0.0075

Salmons and ruffs 23 0.0245 0.0088 0.1050Medium demersal

piscs24 0.0442 0.0035 0.0300 0.0140

Small demersal piscs 25 0.2205 0.1430 0.0807Medium demersal

invert feeders26 0.0050

Small demersal invertfeeders

27 0.0093 0.0031 0.0500 0.0538

Mesopelagics 28 0.0600 0.0600Small demersal

omnivore29 0.0200 0.0150 0.0033 0.0427 0.0001 0.0109 0.0270

Arrow squid 30 0.0245 0.2202Calamary 31 0.0091 0.0112Other squids 32 0.0020 0.0095 0.0050 0.0020 0.0538Octopus 33 0.0115 0.0003 0.0160Large zooplankton

(carnivores)34 0.0296 0.0693 0.1129 0.1310 0.0377 0.0080 0.1684 0.0381 0.0080 0.1000

Small zooplankton(herbivores)

35 0.4938 0.6748 0.0056 0.0347 0.0040 0.9300 0.0024 0.2626 0.8800 0.2000 0.4000 0.4000 0.8000

Benthic grazer(megabenthos)

36 0.5062 0.2883 0.8652 0.0825 0.1036 0.6292 0.0118 0.6602 0.0086 0.0850 0.6992

Detritivore (infauna -macrobenthos)

37 0.1864 0.0815 0.1574 0.9163 0.0479 0.0407 0.5500

Filter feeders 38 0.0049 0.4419 0.0710 0.0018 0.0342 0.0106 0.0012 0.1076Primary production 39 0.0023 0.1348 0.1752 0.5926 0.0039 0.1104 0.8000 1.0000 0.2000Detritus 40Import 41

Page 6: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

S.D. Goldsworthy et al. / Ecological

Fig. 2. Trends in the total landings (catch t y−1) from eight main fisheries in theEGAB ecosystem between 1991 and 2008. (A) All fisheries combined (Total catch),the SASF (Sardine catch) and all fisheries minus sardine catch (Other catch); (B)marine-scalefish line (MS-line); marine-scalefish net (MS-net); Spencer Gulf, GulfSta

lGaBGfifio

tftatl

t Vincent, West Coast prawn fisheries (Prawn); demersal shark (Shark gillnet); GABrawl fishery (GABT); South-east Trawl fishery (SET); and Southern bluefin tuna polend line and purse seine fishery (SBT).

ine, SA Marine Scalefish net, and three prawn fisheries (Spencerulf; Gulf St Vincent; West Coast). Another five fisheries are man-ged by the Australian Government (Commonwealth): Southernluefin Tuna (SBT) purse seine, SBT pole and bait, South East trawl,reat Australian Bight (GAB) trawl, and the gillnet demersal sharkshery (part of the Gillnet Hook and Trap Fishery). Fleet data

rom the SA abalone and southern rock lobster fisheries were notncluded because the EGAB ecosystem model was primarily devel-ped to be a pelagic model.

Retained and discarded catch (landings and discards, respec-ively) data were typically only available for between 1 and 3 yearsor each fishery, and were estimated for 1991 based on their propor-

ion to landed catch (Currie et al., 2009; Fowler et al., 2009; Robertsnd Steer, 2010). All landed and discarded species were assignedheir functional group, and summed (catch) at the functional groupevel (t km−2) (Tables 2 and 3, respectively). Time series of annual

Modelling 255 (2013) 38– 57 43

catch and catch per unit-effort (CPUE) were estimated for 18 and16 functional groups, respectively. A time series of annual fish-ing mortality (F) was used for sardines instead of CPUE becausemarked changes in vessel size in the fleet between 1991 and 2008confounded CPUE estimates (Supplementary Information C).

There were few data on protected species bycatch in mostfisheries; however, estimates were available for common dolphinand Australian sea lion bycatch in the sardine fishery and demersalgillnet shark fishery, respectively (Goldsworthy et al., 2010; Hameret al., 2008). Sea lion bycatch was included as discards, as thefishery overlaps with the foraging distribution of sea lions acrossthe model region (bycatch scaled to 2008 estimated bycatchrelative to historic effort); estimates of dolphin bycatch in thesardine fishery were available between 2005 and 2008 (basedon bycatch rates per unit of fishing effort) (Hamer et al., 2008),and were estimated for the period between 1991 and 2004 usingfishing effort. The sardine fishery impacts dolphins only in part oftheir range, hence bycatch was estimated as fishing mortality (F)by running the model without bycatch mortality, then estimatingF directly from the reconstructed bycatch time-series. The modelwas then re-run including F. Some common dolphin bycatch isalso known to occur in the demersal gillnet shark fishery, but noquantitative data were available.

The mixed trophic impacts routine in the network analysis toolswithin Ecopath was used to evaluate critical trophic interactionsbetween groups in the ecosystem (Supplementary Information, Fig.S1). The routine is based on the method developed by Ulanowiczand Puccia (1990), and allows the computation of direct andindirect impacts which a change in biomass of a predator groupwill have on other groups in the system, assuming that the dietmatrix remains unchanged, and may thus be viewed as a tool forsensitivity analysis.

2.3. Model fitting

Model fitting using the ‘fit to time series’ procedure wasrun in Ecosim using the time-series estimates (1991–2008) ofbiomass and fishing mortality (F) (sardines only), catch and CPUEfor functional groups with available data. Different numbers ofpredator–prey interactions within the dietary matrix were selected(10–90) within this procedure to identify the most sensitive andoptimal number of predator–prey interactions, and their vulner-ability values that would minimise the model sum of squares (SS)and produce the best fit to the time series data (a table with the esti-mated vulnerabilities is provided in Supplementary Information D).

Some of the default Ecosim parameters were then adjusted tofurther decrease the model SS. This included adjusting the max-imum relative feeding time of marine mammals and seabirdsfrom 2.0 (default) to 10.0, and their feeding time adjustmentrates to 0.5 (0 for all other groups), to account for modifica-tions to their search feeding times in response to changes inprey availability (Christensen et al., 2008). Similarly, we adjusteddensity-dependent predator–prey switching power of the dolphinand seal groups from 0 to 2.0, to account for their capacity toopportunistically adjust their diet in response to changes in preyavailability (Piroddi et al., 2010). We also explored improvementsto model fits by adjusting values of density-dependent changes incatchability for pelagic schooling fish such as sardines and tuna(Christensen et al., 2008; Piroddi et al., 2010), but these did notproduce improvements to model fitting.

The final step of the model fitting procedure was to link thebest combination of vulnerability values and the time-series data

to an estimated trend in primary productivity (PP) within theEGAB ecosystem. The importance of seasonal upwelling in theEGAB to enhancing primary, secondary and fish production hasbeen established (Ward et al., 2006). As such, time-series of mean
Page 7: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

44S.D

. G

oldsworthy

et al.

/ Ecological

Modelling

255 (2013) 38– 57

Table 2Summary of fleet landings (catch t km−2) by functional group used in the balanced EGAB Ecopath model. SASF = South Australian sardine fishery.

Group name SASF SAMS-linefisheries

SAMS-netfisheries

SBT purseseine

SBT poleand bait

SE trawl GAB trawl Demersalshark

SG prawnfishery

GSV prawnfishery

WC prawnfishery

Total

Pelagic sharks 0 9.13 × 10−5 8.11 × 10−3 0 0 0 0 1.54 × 10−5 0 0 0 8.22 × 10−3

Demersal sharks 0 1.66 × 10−3 8.21 × 10−3 0 0 2.16 × 10−5 2.26 × 10−5 6.28 × 10−3 0 0 0 1.62 × 10−2

Rays and skates 0 2.29 × 10−4 6.00 × 10−5 0 0 0 2.23 × 10−6 0 0 0 0 2.92 × 10−4

SBT 0 0 0 2.14 × 10−2 9.23 × 10−6 0 0 2.42 × 10−4 0 0 0 2.16 × 10−2

Other tunas-kingfish 0 0 0 1.35 × 10−2 9.17 × 10−5 1.38 × 10−5 1.95 × 10−7 2.51 × 10−4 0 0 0 1.38 × 10−2

Large bentho-pelagicpiscs

0 1.37 × 10−3 4.47 × 10−4 0 0 1.27 × 10−3 3.36 × 10−4 1.00 × 10−7 0 0 0 3.42 × 10−3

Blue mackerel 0 0 0 0 0 0 1.95 × 10−7 0 0 0 0 1.95 × 10−7

Jack mackerel 0 0 0 0 1.03 × 10−6 0 6.49 × 10−8 0 0 0 0 1.09 × 10−6

Sardine 5.72 × 10−5 0 0 0 0 0 0 0 0 0 0 5.72 × 10−5

Salmons and ruffs 0 9.05 × 10−5 4.69 × 10−3 0 0 0 0 5.00 × 10−7 0 0 0 4.78 × 10−3

Medium demersalpiscs

0 2.65 × 10−3 1.73 × 10−3 0 0 5.95 × 10−4 1.05 × 10−3 1.83 × 10−4 0 0 0 6.21 × 10−3

Small demersal piscs 0 0 0 0 0 8.47 × 10−6 0 0 0 0 0 8.47 × 10−6

Medium demersalinvert feeders

0 0 0 0 0 8.47 × 10−6 0 3.30 × 10−6 0 0 0 1.18 × 10−5

Small demersal invertfeeders

0 0 0 0 0 0 0 6.00 × 10−7 0 0 0 6.00 × 10−7

Arrow squid 0 0 0 0 0 5.18 × 10−5 1.09 × 10−5 0 0 0 0 6.27 × 10−5

Calamary 0 9.66 × 10−4 7.49 × 10−4 0 0 0 3.25 × 10−8 0 1.51 × 10−4 1.26 × 10−7 1.48 × 10−10 1.87 × 10−3

Other squids 0 0 0 0 0 0 3.25 × 10−8 0 0 0 0 3.25 × 10−8

Octopus 0 0 0 0 0 0 0 1.00 × 10−7 0 0 0 1.00 × 10−7

Benthic grazer(megabenthos)

0 0 0 0 0 0 1.45 × 10−6 3.00 × 10−7 2.15 × 10−3 1.80 × 10−6 2.11 × 10−9 2.16 × 10−3

Sum 5.72 × 10−5 7.06 × 10−3 2.40 × 10−2 3.48 × 10−2 1.02 × 10−4 1.97 × 10−3 1.42 × 10−3 6.98 × 10−3 2.30 × 10−3 1.93 × 10−6 2.26 × 10−9 7.87 × 10−2

Page 8: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

S.D.

Goldsw

orthy et

al. /

Ecological M

odelling 255 (2013) 38– 57

45

Table 3Summary of fleet discards (catch t km−2) by functional group used in the balanced EGAB Ecopath model. SASF = South Australian sardine fishery.

Group name SASF SAMS-linefisheries

SAMS-netfisheries

SBT purse seine SBT poleand bait

SEtrawl

GABtrawl

Demersalshark

SG prawnfishery

GSV prawnfishery

WC prawnfishery

Total

Aust sea lion 0 0 0 0 0 0 0 2.2 × 10−5 0 0 0 2.17 × 10−5

Pelagic sharks 0 0 0 3.00 × 10−5 0 1.52 × 10−5 0 0 0 0 0 4.49 × 10−5

Demersal sharks 0 5.33 × 10−5 4.63 × 10−6 0 0 0 8.79 × 10−6 4.4 × 10−5 6.46 × 10−4 5.40 × 10−7 6.33 × 10−10 7.57 × 10−4

Rays and skates 0 1.35 × 10−4 2.33 × 10−7 0 0 0 2.09 × 10−5 0 9.22 × 10−4 7.72 × 10−7 9.04 × 10−10 1.08 × 10−3

SBT 0 0 0 0 0 0 0 8.4 × 10−6 0 0 0 8.40 × 10−6

Other tunas-kingfish 0 0 0 0 0 1.40 × 10−6 3.70 × 10−9 1.4 × 10−6 0 0 0 2.80 × 10−6

Large bentho-pelagicpiscs

0 1.34 × 10−3 5.28 × 10−5 0 0 7.84 × 10−5 2.23 × 10−5 0 1.07 × 10−3 8.96 × 10−7 1.05 × 10−9 2.56 × 10−3

Blue mackerel 0 0 0 0 0 0 2.22 × 10−7 0 0 0 0 2.22 × 10−7

Jack mackerel 0 8.20 × 10−6 0 0 0 5.83 × 10−5 3.73 × 10−6 0 1.88 × 10−4 1.57 × 10−7 1.84 × 10−10 2.58 × 10−4

Anchovy 0 0 0 0 0 0 0 0 2.88 × 10−7 2.41 × 10−10 2.82 × 10−13 2.88 × 10−7

Sardine 0 0 0 0 0 0 0 0 3.79 × 10−7 3.17 × 10−10 3.71 × 10−13 3.79 × 10−7

Inshore smallplanktivores

0 0 0 0 0 0 0 0 3.09 × 10−7 2.58 × 10−10 3.02 × 10−13 3.09 × 10−7

Salmons and ruffs 0 0 4.84 × 10−5 0 0 0 0 0 6.64 × 10−8 5.56 × 10−11 6.51 × 10−14 4.85 × 10−5

Medium demersalpiscs

0 2.05 × 10−4 1.11 × 10−3 0 0 0 6.39 × 10−5 4.0 × 10−7 7.08 × 10−4 5.93 × 10−7 6.94 × 10−10 2.08 × 10−3

Small demersal piscs 0 0 1.14 × 10−5 0 0 0 0 0 3.10 × 10−3 2.60 × 10−6 3.04 × 10−9 3.12 × 10−3

Medium demersalinvert feeders

0 1.23 × 10−5 1.01 × 10−6 0 0 6.04 × 10−6 0 1.0 × 10−7 1.74 × 10−3 1.461 × 10−7 1.71 × 10−10 1.94 × 10−5

Small demersal invertfeeders

0 0 0 0 0 0 0 0 1.99 × 10−4 1.66 × 10−7 1.95 × 10−10 1.99 × 10−4

Small demersalomnivore

0 0 1.61 × 10−3 0 0 0 5.20 × 10−6 0 3.64 × 10−3 3.05 × 10−7 3.57 × 10−10 1.98 × 10−3

Arrow squid 0 0 0 0 0 0 4.10 × 10−7 0 0 0 0 4.10 × 10−7

Calamary 0 0 0 0 0 0 0 0 4.68 × 10−4 3.92 × 10−7 4.59 × 10−10 4.69 × 10−4

Benthic grazer(megabenthos)

0 0 1.24 × 10−4 0 0 0 0 0 2.08 × 10−3 1.74 × 10−6 2.04 × 10−9 2.20 × 10−3

Filter feeders 0 0 0 0 0 0 1.13 × 10−5 0 1.00 × 10−4 8.37 × 10−8 9.80 × 10−11 1.11 × 10−4

Sum 0 1.75 × 10−3 2.96 × 10−3 3.00 × 10−5 0 1.59 × 10−4 1.37 × 10−4 7.6 × 10−5 1.00 × 10−2 8.39 × 10−6 9.82 × 10−9 1.51 × 10−2

Page 9: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

4 ogical

mce(2rwcpMi

2

emsgPlatctos

TBQ

6 S.D. Goldsworthy et al. / Ecol

onthly upwelling anomalies, calculated from the alongshoreomponent of wind stress at South Neptune Island (Middletont al., 2007), were considered as potential PP forcing functionsFF) in the EGAB ecosystem. Wind stress values between 1991 and008 were scaled to positive values, and then converted to valueselative to that of January 1991. Using the anomaly search functionithin Ecosim, we set the PP variance to 50 to enable the model to

apture changes in the magnitude of ‘spikes’. The number of splineoints was set to 18, to track annual anomalies in the time series.odel-generated time series were compared to the upwelling

ndex time series for agreement.

.4. Ecosystem indicators

After the model fitting procedure in Ecosim, we examined fourcosystem indictors that can be used to evaluate changes in thearine ecosystem. (1) total catch; (2) Kempton’s index of biodiver-

ity (Q), which expresses biomass species diversity of functionalroups with a trophic level (TL) of three or higher (Ainsworth anditcher, 2006; Kempton and Taylor, 1976); (3) the mean trophicevel of the catch (mTLC) which is calculated as the weighted aver-ge of the TL of fishery targeted species (Pauly et al., 1998); and (4)he Fishing in Balance Index (FIB index), which assesses whether

atch rates are in balance with ecosystem trophic production dueo catch at a given TL being related to the assimilation efficiencyf the ecosystem (Coll et al., 2009). The FIB index will remain con-tant if a decline in mTLC is matched by an ecologically appropriate

able 4iological parameters by functional group of the balanced EGAB Ecopath model.

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

Group name Trophic level B

1 Baleen whales 3.01

2 Bottlenose dolphin 4.61

3 Common dolphin 4.66

4 NZ fur seal 4.80

5 Aust fur seal 4.53

6 Aust sea lion 4.91

7 Little penguin 4.71

8 Petrels 4.09

9 Gannets 5.01

10 Terns 4.52

11 Pelagic sharks 4.92

12 Demersal sharks 3.92

13 Rays and skates 3.68

14 SBT 4.52

15 Other tunas-kingfish 4.50

16 Large bentho-pelagic piscs 4.68

17 Blue mackerel 3.23

18 Jack mackerel 3.22

19 Redbait 3.38

20 Anchovy 3.63

21 Sardine 3.36

22 Inshore small planktivores 3.93

23 Salmons and ruffs 4.51

24 Medium demersal piscs 3.47

25 Small demersal piscs 2.66

26 Medium demersal invert feeders 4.00

27 Small demersal invert feeders 3.53

28 Mesopelagics 3.07

29 Small demersal omnivore 3.77

30 Arrow squid 4.18

31 Calamary 4.50

32 Other squids 3.14

33 Octopus 4.16

34 Large zooplankton (carnivores) 2.20

35 Small zooplankton (herbivores) 2.00 336 Benthic grazer (megabenthos) 3.24 137 Detritivore (infauna - macrobenthos) 2.52 338 Filter feeders 2.80

39 Primary production 1.00 140 Detritus 1.00 1

Modelling 255 (2013) 38– 57

increase in catch, and conversely for increasing trophic level (Paulyand Palomares, 2005). In general, the index increases if the underly-ing fishery expands beyond its traditional fishing area or ecosystem,and decreases if the geographic area contracts, or if the underlyingfood web is collapsing (Pauly and Palomares, 2005).

2.5. Model scenarios

We used the fitted Ecosim EGAB model to explore thespecies/group responses (increase/decrease in biomass) to reduc-tions in sardine biomass (to 0.75, 0.50 and 0.25 of 2008 biomass).The fishery effort time series was extended to year 2040, main-taining the fishing effort of all fleets except sardine beyond 2008at 2008 levels, and increasing the sardine fishery effort to drivereductions in sardine biomass. The time series of sardine fishingmortality was deselected in these procedures. In these scenarios,both NZ and Australian fur seal biomass was forced to increasebetween 2008 and 2040 (in line with observed increases between1991 and 2008), so that population abundances in the EGAB wereabout 1.5 and 7.0 times present levels (2008), respectively. This isequivalent to an increase from ∼73,000 to ∼111,000 for NZ fur sealsand ∼2800 to ∼20,000 for Australian fur seals between 2008 and2040. In addition, simulations with and without fishing effort onSBT were performed by adjusting effort in the purse seine fishery

beyond 2008. Scenarios where arrow squid biomass was reducedby 50% and 75% were also examined. For all scenarios, we comparedthe relative change in biomass between 2008 and 2040 with andwithout PP forcing.

Parameters in bold were estimated by the model. P/B = production/biomass;

iomass (t km−2) P/B (year−1) Q/B (year−1) EE

0.0389 0.020 5.097 0.0000.00611 0.080 16.566 0.0000.0039 0.090 20.511 0.0000.00453 1.184 47.526 0.9440.00047 1.157 28.819 0.9830.00422 0.792 29.445 0.1500.000698 1.250 85.600 0.9940.00306 1.000 147.100 0.0700.0000308 1.000 138.300 0.0000.00000635 1.000 89.900 0.0000.0459 0.200 1.200 0.9000.307 0.180 1.800 0.3260.459 0.350 2.700 0.0140.145 0.200 1.600 0.9000.0769 0.200 1.200 0.9000.452 0.338 3.315 0.9590.219 0.370 3.500 0.8570.919 0.470 3.300 0.9000.561 0.740 2.800 0.9001.272 0.700 5.040 0.5551.517 1.600 5.040 0.3300.489 1.010 7.300 0.9000.246 0.440 5.400 0.9000.302 0.485 5.400 0.8441.467 0.853 5.367 0.4120.0786 0.860 5.400 0.9600.149 1.090 5.500 0.9000.106 1.005 6.673 0.9000.170 0.840 16.000 0.9570.341 1.950 3.900 0.9000.0837 1.950 3.900 0.9000.111 2.500 5.850 0.9000.294 2.500 5.850 0.9001.287 20.000 70.000 0.8005.119 5.000 32.000 0.8001.013 1.600 6.000 0.8000.983 1.600 6.000 0.8001.581 1.600 6.000 0.8004.900 745.000 0.000 0.1080.0 0.009

Page 10: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

ogical

3

3

wiwpbagdbtdomfs

ppfistfiniaapnt

FbS

S.D. Goldsworthy et al. / Ecol

. Results

.1. Trophic structure and flow

The balanced model parameters of the 40 functional groupsithin the Ecopath model are presented in Table 4. The balanc-

ng procedure required adjustment to the diets of some groupshere ecotrophic efficiencies (EE) were initially >1. EE is the pro-ortion of production that is either harvested or predated upony higher trophic levels and cannot exceed 1. The main dietarydjustments included reduction in the contribution of little pen-uins and petrels in the diets of NZ fur seals, and of petrels in theiets of Australian sea lions; reduction in the contribution of largeentho-pelagic fish in the diet of arrow squid; reduction in the con-ribution of medium demersal invertebrate feeders in the diets ofemersal sharks and octopus; and reductions in the contributionf small demersal omnivores in the diets of salmon and ruffs and ofedium demersal piscivorous fish. The trophic flows between the

unctional groups in the EGAB ecosystem estimated by Ecopath, areummarised in Fig. 3.

Results from the mixed trophic impacts routine indicate theositive effect of primary production on small and large zoo-lankton, the positive effects of these groups on small pelagicshes which in turn have a positive effect on dolphin, seal andeabird groups (Supplementary Information, Fig. S1). Groups con-aining commercially targeted species influenced their respectiveshing fleets positively, and most groups affected themselvesegatively (Fig. S1). Cephalopod groups affected small demersal

nvertebrate feeders and meso-pelagic fish negatively; salmonsnd ruffs affected medium and small demersal piscivorous fish

nd inshore small planktivores negatively; and bentho-pelagiciscivorous fish affected small pelagic fish and octopus groupsegatively (Fig. S1). SBT affected other tunas and kingfish nega-ively; demersal sharks affected medium demersal invertebrate

ig. 3. Food web diagram of the Ecopath model of the EGAB ecosystem. Functional groupetween groups are indicated by lines. Ovals (dashed) encircle major trophic groups inclumall Pelagic fish group are encircled by a solid oval.

Modelling 255 (2013) 38– 57 47

feeders, pelagic sharks and sea lions negatively; and sea lionsaffected rays and skates and little penguins negatively; NZ furseals affected little penguins and petrels negatively (Fig. S1). Of thefisheries, the SA marine scalefish net fishery affected pelagic anddemersal sharks negatively; the SBT purse seine fishery affectedSBT, other tuna and kingfish negatively; the demersal shark fisheryaffected sea lions and demersal sharks negatively, and the SpencerGulf prawn fishery affected rays and skates negatively (Fig. S1).Based on this analysis, the sardine fishery has negligible impactson other groups. However, as a steady-state analysis, all theseassessments do not take into account the changing abundances ordiets of groups. Such dynamics are explored in Ecosim scenariosbelow.

The proportional breakdown of the consumption by predators ofsardines and of all combined small pelagic fish groups is presentedin Fig. 4. The key predators of sardines in ranked order were largebentho-pelagic fish, arrow squid, salmons and ruffs, SBT, other tunaand kingfish, common dolphins and pelagic sharks. For all smallpelagic fish, the key predators in ranked order were salmons andruffs, arrow squid, SBT, calamary, other tuna and kingfish, commondolphins, little penguins, NZ fur seals, petrels and pelagic sharks(Fig. 4). The total consumption of sardines estimated by the Ecosimmodel was 123,369 t y−1, representing about a third (31%) of thetotal combined consumption of small pelagic fish (396,129 t y−1) inthe EGAB ecosystem.

3.2. Time-series fitting

Ecosim model fits to fishery time series and model-derivedwind-stress anomaly forcing functions (Fig. S2) are presented in

Fig. 5 There was a high degree of variability in the degree to whichEcosim reproduced the CPUE (biomass) and yield (catch) trendsof various functional groups (Fig. 5). Trends in modelled biomassfitted observed trends reasonably well for pelagic sharks, demersal

s are indicated by circles, with size proportional to their biomass. The trophic linksding marine mammals, seabirds, pelagic sharks, large pelagic fish and cephalopods.

Page 11: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

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 andi cosyst1

sdcdbdsebawc(guuiitfitm

nshore planktivores) by their predators based on the Ecopath model of the EGAB e23,369 t y−1 and 396,129 t y−1, respectively.

harks, large bentho-pelagic piscivorous fish, sardine, mediumemersal piscivorous fish, and calamary (Fig. 5). Trends in modelledatch fitted observed trends reasonably well for pelagic sharks,emersal sharks, rays and skates, SBT, other tuna and kingfish, largeentho-pelagic piscivorous fish, jack mackerel, sardine, mediumemersal piscivorous fish, small demersal piscivorous fish, arrowquid and benthic grazers (Fig. 5). In many instances the fit of mod-lled trends to observed trends was substantially better for eitheriomass or catch. For example for SBT, jack mackerel, sardine andrrow squid, the fit to catch was much better than the fit to biomass;hereas for pelagic sharks, medium demersal piscivorous fish and

alamary, the fit to biomass was much better than the fit to catchFig. 5). The latter instances were predominantly for functionalroups caught across multiple fishing fleets, and because Ecosimses only one fleet’s CPUE time series to estimate biomass (butses all fleet effort data to estimate catch), such variability in fits

s not surprising. For SBT, the capacity for Ecosim to fit to changesn biomass was confounded by the transition from pole and bait as

he main fishing method to purse seining, with the pole and baitshery ending in 2000. CPUE from the purse seine fishery was usedo drive biomass trends for this group. The congruence between

odelled and observed trends in biomass and catch was best for

em. Total consumption of sardines and of all small pelagic fishes is estimated to be

functional groups composed of species targeted by fisheries, suchas demersal sharks (gummy and school shark), SBT, large bentho-pelagic fish (snapper), sardine, medium demersal piscivorous fish(King George whiting), calamary and benthic grazers (western kingprawn); and poorer for non-targeted (byproduct) species such asblue and jack mackerel and small demersal piscivorous fish (Fig. 5).

A summary of the estimated changes in the biomass of theEGAB ecosystem between 1991 and 2008 summarised into ninebasic groups, with and without PP forcing in the Ecosim model,is presented Table 5. With no change in PP, the observed increasein high trophic predators (fur seals) over the period occurs to thedetriment of large fish (SBT and other tuna and kingfish mainlydue to over fishing) and of medium and small fish. However, theCPUE/biomass trends for many functional groups (pelagic shark,demersal shark, large bentho-pelagic fish, sardine, salmons andruffs, medium demersal piscivorous fish, arrow squid, calamary andbenthic grazers, Fig. 5) suggest that changes in fishing effort aloneare not enough to capture the observed biomass trends in these

groups, and that the best model fits with minimum SS resultedfrom PP forcing, suggesting that an overall increase in PP is alsorequired to support the observed biomass increases in many of thegroups between 1991 and 2008 (P < 0.0001, Table 5 and Fig. 5).
Page 12: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

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

F UE) at of var

3

wiratTsect(

t

TC(

ig. 5. Time series fits of the EGAB Ecosim model (thin line) to observed biomass (CPrend lines (dashed) are provided, together with values of the slope and coefficient

.3. Temporal changes in group biomass

The sardine fishery grew significantly between 1991 and 2008,ith the major growth period occurring since 2000 (Fig. 2). Land-

ngs of SBT increased by 75% between 1991 and 2008 (linearegression, P = 0.007), while landings from the marine scalefish netnd demersal shark fisheries declined by 88% and 35%, respec-ively (linear regression, P < 0.001 and P = 0.0015, respectively).here has been little change in the landings from the marinecalefish line fisheries (+7%, linear regression, P = 0.1750) and west-rn king prawn (−1%, linear regression, P = 0.741, three regionsombined); although GAB trawl fishery landings declined by 35%,

he decrease was not significant (linear regression, P = 0.085)Fig. 2).

The most significant increases in population sizes of apex preda-ors between 1991 and 2008 were for NZ fur seals, Australian

able 5hange of estimated relative biomass of key taxa groups based on Ecosim simulations of tPP) forcing; and between 2008 and 2040 with no PP forcing, with SBT fishery catch remo

Summary groups Relative biomass change of groups

1991–2008 1991–2008

No PP forcing PP forcing

Marine mammals, birds, pelagic sharks 54% 87%

Demersal sharks and rays 4% 35%

Large piscivorous fish −13% 38%

Cephalopods 4% 95%

Medium-small fish −4% 78%

Small and meso-pelagic fish 25% 25%

Benthic grazers/filter feeders/detritovores 0% 30%

Zooplankton 0% 37%

Primary production 0% 16%

nd catch data (dots) for 16 functional groups between 1991 and 2008. The modellediation (r2) (see overleaf for remaining plots).

fur seals, Australian sea lions, and bottlenose dolphins (linearregression, P < 0.0001) but there were no significant changes inthe biomass of little penguins and petrels (Fig. 6A). Pelagic anddemersal shark biomass increased significantly (linear regression,P < 0.0001) when fishing mortality was reduced as a result of reduc-tions in effort in the marine scalefish net fishery and the demersalshark fishery (Fig. 6B). Rays and skates also increased but pre-dicted trends are at odds with CPUE data, especially between 2002and 2008 (Fig. 5). Biomass of large bentho-pelagic piscivorousfish was also predicted to increase (linear regression, P < 0.0001),while the biomass of SBT and other tuna and kingfish groupsdeclined (linear regression, P < 0.0001), principally in response to

high fishing mortality. All the small pelagic fish groups (bluemackerel, jack mackerel, redbait, anchovy, sardine, inshore smallplanktivores, salmon and ruffs) and squids (arrow squid, calamary)were estimated to increase in biomass between 1991 and 2008

he EGAB ecosystem between 1991 and 2008, without and with primary productionved and with a 50% reduction in arrow squid biomass.

2008–2040 2008–2040 2008–2040No PP forcing No PP forcing

No SBT fisheryNo PP forcingNo SBT fishery<50% squid biomass

62% 85% 80%3% 3% 4%2% 90% 96%

−7% −7% −23%1% 1% 1%

−27% −27% −3%1% 1% 1%1% 1% 1%0% 0% 0%

Page 13: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

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

Fig. 6. Estimated changes in the biomass of Ecosim EGAB model groups between1991 and 2008: (A) marine mammals and birds (biomass change in NZ and Aus-tralian fur seals is forced in the model as these estimates are based on empiricaldata for these species in the EGAB ecosystem); (B) sharks, rays and skates, tunas andkingfish; and (C) small pelagic fish, arrow squid and calamary.

Fig. 7. Ecosystem indicators calculated from the Ecosim EGAB model for the period1991 to 2008. (A) Kempton’s Q biomass diversity index; (B) Mean trophic level of thecatch and (C) Fishing in balance (FIB) index. Estimated trends are given by dashedlines (regression), with their coefficient of variation (r2) and level of significance (P).

Page 14: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

ogical

(eri2

3

Eotorisf

Fb2t

S.D. Goldsworthy et al. / Ecol

linear regression, P < 0.0001) (Fig. 6C). Common dolphins werestimated to have declined between 1991 and 2004, with someecovery occurring since 2004 when management measures werentroduced to mitigate bycatch in sardine fishery (Hamer et al.,008).

.4. Ecosystem indicators

The ecosystem indicators identified significant changes in theGAB ecosystem between 1991 and 2008 (Fig. 7A–C). Total catchf the combined fisheries showed a 4.6-fold increase, attributedo growth in the sardine fishery (Fig. 2A). The total catch of allther fisheries combined showed a slight (0.89) but non-significant

eduction over the same period (Fig. 2A). Kempton’s Q biodiversityndex usually increases with growing biomass of high trophic levelpecies, and decreases with increased fishing impacts. The indexor the EGAB ecosystem increased significantly between 1991 and

ig. 8. Predicted change in the biomass of functional groups in the Ecosim EGAB model iniomass. In this simulation, NZ fur seal and Australian fur seal groups were forced to inc008), so that population abundances in the EGAB were about 1.5 and 7.0 times present

he scale of the x-axis.

Modelling 255 (2013) 38– 57 51

2008 (Fig. 7A), commensurate with increasing biomasses of marinemammals, sharks, and some large piscivorous fish and squid (Fig. 6).The mTLC decreased significantly between 1991 and 2008 (linearregression, P < 0.001) (Fig. 7B), as a consequence of growth in thesmall pelagic fishery for sardines. The fishing in balance (FIB) indexbetween 1991 and 2008 generally increased but not significantlyand all values were <0.02 (Fig. 7C). There was a temporary spike inFIB in 2005 that corresponds to the peak in sardine catch in thatyear (40,000 t, Fig. 2A), but it then declined back to similar levelsto between 2002 and 2004 (Fig. 7C). An FIB index <0 occurs wherefishing impacts are so high that the ecosystem function is impaired,or where discarding occurs and it is not considered in the analysis(Christensen, 2000). An FIB index = 0 indicates high production at

2040 relative to 2008 with sardine biomass reduced to 0.75, 0.50 and 0.25 of 2008rease between 2008 and 2040 (in line with observed increases between 1991 and(2008) levels, respectively (with no change in sardine biomass). Note: the break in

lower trophic levels with fishing in balance; and an FIB index >0indicates an expansion of fishing and/or where bottom-up effectsare occurring, resulting in more catch than expected (Coll et al.,2009).

Page 15: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

5 ogical

3

adfisa(lsfa

iba

Fwl2

2 S.D. Goldsworthy et al. / Ecol

.5. Model scenario results

Functional groups that declined below their 2008 biomasss a result of reductions in sardine biomass included: commonolphins, little penguins, gannets, terns, SBT, other tunas and king-sh, small demersal invertebrate feeders, mesopelagics, and arrowquid (Fig. 8). For some declining groups, their decline was lesseneds fishing on sardines increased (e.g. petrels and jack mackerel)Fig. 8). For some groups with increasing biomass, the increase wasessened (e.g. NZ fur seals, Australian sea lions, pelagic sharks, andalmon and ruffs), or enhanced (e.g. bottlenose dolphins, Australianur seals, redbait, anchovy, small demersal omnivores, other squidsnd octopus) as sardine biomass was reduced (Fig. 8).

Of the land-breeding marine predators, those most directly

mpacted by reductions in sardine biomass were terns, whoseiomass declined by 9%, 20% and 35% in response to 25%, 50%nd 75% reductions in sardine biomass, respectively; and gannets

ig. 9. Predicted change in the biomass of functional groups in the Ecosim EGAB modelere simulated by reducing effort in the purse seine fishery by adjusting time series value

evels; 50% 2008 levels; 25% 2008 levels and no fishing effort). In these simulations, NZ

040 (in line with observed increases between 1991 and 2008), so that populations abun

Modelling 255 (2013) 38– 57

whose biomass declined by 6%, 17%, 27%, respectively (Fig. 8).Little penguin decline was less affected by sardine reductions withdeclines of 12.0%, 12.5% and 12.9% in response to 25%, 50% and 75%reductions in sardine biomass, respectively. Relative biomasses ofNZ fur seals (65%, 61%, and 56%) and Australian sea lions (28%, 20%and 14%), were lower, respectively in response to 25%, 50% and 75%reductions in sardine biomass, due mainly to indirect interactionsreducing the biomass of other key prey taxa (arrow squid and largebentho-pelagic piscivorous fish, respectively) (Fig. 8).

Our simulations of continuation of the current (2008) purseseine fishing effort on tuna beyond 2008 resulted in a 21% declinein biomass by 2040, on top of the estimated 51% decline up to 2008.Simulations showed SBT biomass increased when fishing effortwas reduced by 50%, 75% and 100% from 2008 levels, resulting

in increases in SBT biomass of 45%, 148% and 615%, respectivelybetween 2008 and 2040 (Fig. 9). They also resulted in markedincreases in the biomass of the other tuna and kingfish group (58%,

in 2040 relative to 2008 with biomass increases of SBT. Biomass increases in SBTs between 2008 and 2040 under different scenarios of fishing effort (current [2008]fur seal and Australian fur seal groups were forced to increase between 2008 anddances in the EGAB were about 1.5 and 7.0 times present (2008), respectively.

Page 16: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

ogical

1iptws

w(fim(etc

Fa(ft

S.D. Goldsworthy et al. / Ecol

99% and 1159%, respectively). Recovery of SBT was coincident withncreases in the biomass of pelagic sharks, but does not have majorositive or negative effects on any other groups (Fig. 9). Althoughhe consequences are not examined here, the global quota of SBTas coincidentally cut by 20 per cent for the 2010 and 2011 fishing

easons since completion of this study.Simulations with and without fishing effort on SBT, and with and

ithout PP forcing indicated that the recovery of apex predatorsmarine mammals, birds and pelagic sharks) and of large pelagicsh such as SBT is supported through consumptions of low andid-tropic species, especially small pelagic fish and cephalopods

Table 5). The main small pelagic fish group impacted is jack mack-

rel, with negligible change in other small pelagic species, whilehe most negatively impacted cephalopods include arrow squid andalamary (Fig. 10).

ig. 10. Predicted change in the biomass of functional groups in the Ecosim EGAB model

nd with no SBT fishery. Biomass reduction in arrow squid were simulated by adjusting ticurrent [2008] levels; 50% biomass reduction of 2008 levels; 75% biomass reduction of

orced to increase between 2008 and 2040 (in line with observed increases between 199imes present (2008), respectively.

Modelling 255 (2013) 38– 57 53

The potential role of reduced competition for small pelagic fishby a reduction in biomass of arrow squid was explored by gradu-ally increasing fishing mortality in the time series between 2008and 2040, while excluding SBT fishing effort and enabling fur sealsto recover as above. Results from scenarios where arrow squidbiomass was reduced by 50% and 75% are presented in Fig. 10, andrelative changes in the biomass of the EGAB ecosystem summari-sed into nine basic groups between 2008 and 2040 are presentedin Table 5. Results indicate that for some high trophic-level preda-tor groups that increased in biomass over the 2008–2040 periodand were important predators of arrow squid (NZ fur seals, pelagicsharks, SBT, and other tunas and kingfish), rates of increase were

reduced commensurate with reductions in arrow squid biomass.However, most others, including common dolphins, Australianfur seals, Australian sea lions, little penguins, petrels, gannets,

in 2040 relative to 2008 with different scenarios of arrow squid biomass reductionme series values between 2008 and 2040 under different scenarios of fishing effort2008 levels). In these simulations, NZ fur seal and Australian fur seal groups were1 and 2008), so that populations abundances in the EGAB were about 1.5 and 7.0

Page 17: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

5 ogical

taecrtfiwssabpfi

4

totibmisabotes(drettmmitsertto

kitfitbdspo

1fiiis

4 S.D. Goldsworthy et al. / Ecol

erns, and demersal sharks, responded positively to reductions inrrow squid biomass (Fig. 10). Small pelagic fish species (with thexception of inshore small planktivores), large bentho-pelagic pis-ivorous fish and salmon and ruffs also responded positively toeductions in arrow squid biomass (Fig. 10). Comparison betweenhe scenarios of increasing fur seal biomass with removal of SBTshing effort, and the same scenario where arrow squid biomassas reduced by 50% (last two columns of Table 5) highlight the

ignificant predation pressure that cephalopods (especially arrowquid) probably place on small pelagic fish, and how with reducedrrow squid biomass, greater production of small pelagic fish cane directed into higher trophic levels (in this instance, the increasedroduction of small pelagic fish was directed into large piscivoroussh and other apex predator species (Table 5)).

. Discussion

This study presents the first trophic model of the EGAB ecosys-em. Ecopath with Ecosim was used to resolve the trophodynamicsf the EGAB ecosystem, to improve our understanding of therophic interactions of the major functional groups and to assess thempacts of temporal changes in fishing activity and apex predatoriomass between 1991 and 2008. The model results suggest thatost apex predator populations have increased over this period,

ncluding baleen whales, bottlenose dolphins, fur seals, pelagicharks, large bentho-pelagic piscivorous fish, small pelagic fishnd cephalopods. Fur seal recovery is likely to have been driveny a demographic response to reduced mortality from the 1970snwards that enabled remnant populations to overcome popula-ion reductions from colonial sealing that occurred 140–170 yearsarlier (Ling, 1999). In contrast, the recovery of pelagic and demer-al sharks, large bentho-pelagic piscivorous fish and some othernon-tuna) fishes indicated by the model, appears primarily to beriven by reductions in fishing mortality. This may partially beelated to reductions in domestic tuna fishing TACs since polingnded, and the expulsion of the Japanese/international longlininguna fleets from within Australia’s EEZ including the GAB wherehey fished through the late 80s and early 90s. However, the Ecosim

odels suggest that positive PP forcing over the period improvesodel fits, and therefore may have underpinned uniform increases

n small pelagic fish and cephalopod biomass that have contributedo the broad scale increase in these higher trophic levels (Fig. 9). Atudy by Brown et al. (2010) simulated positive increases in PP toxamine the ecosystem responses to climate change scenarios in aange of marine ecosystems in Australia. As with the EGAB model,hey found that such increases in PP also generally led to a posi-ive increase to most functional groups and to increased biomassf high trophic species (Brown et al., 2010).

In contrast, the projected decline in SBT and other tunas andingfish biomass in the EGAB ecosystem appear to be driven byncreases in fishing effort. Southern bluefin tuna are a highly migra-ory species and the purse seine fishery in the EGAB targets juvenilesh in the 3–5-year-old age classes (Wilson et al., 2009). The extento which modelled changes in the EGAB reflect changes in theroader population, or are compounded by impacts in other juris-ictions and fisheries (High Seas long-line fisheries) is unclear, buthould be investigated further. For example, the pelagic sharks andetrels cross several management jurisdictions within time-scalesf <1 y.

The growth of the sardine fishery since its establishment in991 has been rapid, and its catch now exceeds that of all other

sheries in the region by a factor of three. Despite this, sensitiv-

ty analyses based on mixed trophic impacts detected negligiblempacts of the fishery on other predators. However, this steady-tate analysis was based on the establishment years of the fishery

Modelling 255 (2013) 38– 57

when catches were very low, and does not take into account thesubstantial increases in catch and the changing abundances ordiets of groups. In contrast, dynamic assessment using Ecosimindicated that many groups were sensitive to changes in sardinebiomass.

Of the land breeding marine predators, crested tern popula-tions demonstrated the greatest sensitivity to reduction in sardinebiomass, both in direction (negative) and magnitude, followed byAustralasian gannets. The latter species only breeds at one site offCape Jaffa some distance from the centre of the sardine fishery. Incontrast, there are at least 10 breeding colonies of crested ternssituated adjacent to the sardine fishery in southern Spencer Gulfand Investigator Strait (McLeay et al., 2010). Demographic stud-ies of crested terns found that they were physically smaller andhad lower survival rates in years following the two sardine massmortality events, in 1995 and 1998 (McLeay et al., 2009b). Theirdependency on small pelagic fish, including sardines which theytake from waters near colonies during a short breeding season(McLeay et al., 2009a, 2010), gives this species the greatest potentialto provide ecological performance indices for the sardine fishery,through foraging, reproductive and population response variables;and further research to assess their suitability is warranted.

Little penguins also demonstrated a slight reduction in biomassin response to reduced sardine biomass, although the magnitudeof this response was small. In Victoria, the reproductive success oflittle penguins decreased coincident with a reduction in the con-tribution of sardines in their diet due to a sardine mortality event(Chiaradia et al., 2003; Dann et al., 2000). The Ecosim model identi-fied that the biomass increase of New Zealand fur seals reduced inresponse to a reduction in sardine biomass. The Ecosim model alsoidentified common dolphins, pelagic sharks, SBT, and other tunaand kingfish as sensitive to biomass reductions in sardines, but assome are subject to their own fishery related mortality, are lesssuitable as indicator taxa compared to the land breeding marinepredators.

Ecosystem indicators identified significant changes in the EGABecosystem between 1991 and 2008, the most significant of whichwas the 4.6-fold increase in total catch, which was entirelyattributable to growth in the sardine fishery within southernSpencer Gulf and the Investigator Strait (Figs. 2A and 7A–C). Notsurprisingly, this led to a significant reduction in the mTLC between1991 and 2008. Usually such reductions demonstrate increasedecosystem impact as a consequence of a reduction in high trophiclevel relative to lower trophic level organisms (Coll et al., 2009).However, the significant increase in Kempton’s Q biodiversity indexover the period indicates increased growth in high trophic levelspecies, suggesting that reduction in the mTLC is attributable togrowth in the sardine fishery, and not a reduction in high trophiclevel biomass. Furthermore, FIB index values generally close to zerosuggest high production at low trophic levels with fishing in bal-ance. The increasing trend in FIB index over the study period, mayreflect geographic expansion of fishing effort and/or more catchthan expected due to bottom-up effects such as increased primaryproductivity (Coll et al., 2009). Values above zero indicate fishingin balance, whereas values below zero suggest high fishing impactsthat are impairing ecosystem function (Christensen, 2000). The FIBindex was developed in part to address the situation where a reduc-tion in the mTLC occurs as a consequence of deliberate targeting oflower trophic level species, rather than due to fishing down marinefood webs. Another potential impact when a low trophic level fish-ery expands rapidly is the replacement of predation mortality withfishing mortality, while overall total mortality remains relative con-

stant. Ecosim results for the sardine fishery suggest this has notoccurred in the EGAB ecosystem. Instead predation mortality hasincreased as fishing mortality and total mortality has increased(Supplementary Information, Fig. S3), with predation mortality as a
Page 18: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

ogical

pa

gtasmhtsbatdei(fiteac

aM2esSttObshghact

rsocapehprapbamgostetbct

S.D. Goldsworthy et al. / Ecol

roportion of total mortality increasing significantly between 1991nd 2008 (linear regression, P < 0.001, r2 = 0.616).

In combination, the above indices are useful because they inte-rate many aspects of fishing impacts and trophodynamic changeo provide measures of ecosystem health. For the EGAB ecosystem,ll these indices suggest that despite the large scale expansion of theardine fishery between 1991 and 2008 and reduction in the mTLC,ost functional groups from cephalopods to small pelagic fish and

igh trophic level predators have been increasing in biomass overhis period. This suggests that the current fishery managementtrategy is sufficiently conservative to ensure that the fishery iseing managed according to the principles of ecological sustain-ble development (ESD) (Fletcher et al., 2002). However, the extento which this positive assessment reflects management of the sar-ine fishery is uncertain, given that much of the positive changesstimated by the model reflect other changes in the ecosystem,ncluding reductions in fishing effort and mortality in some fleetsSA marine scalefish net and Commonwealth demersal gillnet sharkshery), and a positive trend in PP over the study period. Howhe current or alternate management strategies of the sardine fish-ry would perform under alternate oceanographic conditions (suchs more variable upwelling), and/or in response to managementhanges in other fisheries in the region, may be useful to explore.

One of the key dynamics in the EGAB ecosystem is the vari-ble trends in biomass of some of the high trophic level species.ost notable is the recovery of fur seal populations over the last

5 years to numbers not seen for more than 150 years. This recov-ry is continuing and it is unclear at what level populations willtabilise. Fishing pressure on populations of SBT in the GAB andouthern Ocean south of Australia over the last 30 years has seenhese stocks reduced to a fraction of their pre-exploited levels, andhey are subject to continued over fishing (Wilson et al., 2009).ther important pelagic predators, such as pelagic sharks have alsoeen targeted by some fisheries and are subject to bycatch, and haveeen historic declines. How such depletions in populations of keyigh-trophic level predators over the last 200 years (a ‘predator-ap’) has affected the EGAB ecosystem is unclear, but this studyas provided some interesting insights, in particular hypothesesbout the dynamics between cephalopods and their predators andompetitors, and the importance of these interactions in regulatinghe flow of small pelagic fish production to high trophic levels.

Ecosim scenarios that explored the ecosystem response toecovering fur seal and SBT populations out to 2040, indicated thatuch recoveries were supported through reductions in the biomassf small pelagic fish (mainly jack mackerel) and cephalopods (espe-ially arrow squid). Scenarios investigating the response to reducedrrow squid biomass as a consequence of predation and com-etition pressure exerted by apex predators, suggest this maynable greater production of small pelagic fish to be directed intoigher trophic levels, in this case to large piscivorous fish and apexredator species. This hypothesis suggests that ‘predator gaps’ thatesulted from reduced biomass of long-lived predator species suchs fur seal, SBT and shark, may have been quickly filled by semel-arous species, such as cephalopods (squids) which can build upiomass quickly in response to increased small pelagic fish avail-bility and reduced predation pressure. A build-up in cephalopodsay reduce trophic flows to high trophic levels, with their short

eneration times and reduced predation pressure resulting in muchf their biomass being returned to detritus. A study of the role ofmall pelagic fish in southern Australian ecosystems identified thathese ecosystems were largely bottom-up forced, but that differ-nt parts of the these food webs could exert both bottom-up and

op-down control (Bulman et al., 2010). Furthermore, switchingetween these states may occur in response to fishing and climatehange pressures (Bulman et al., 2010). In this respect, it is possiblehat cephalopods may be able to exert some degree of ‘wasp-waist’

Modelling 255 (2013) 38– 57 55

control (Cury et al., 2000), by restricting energy flow to competi-tors at similar or higher trophic levels, although typical wasp-waistspecies are highly mobile low-trophic level species, such as smallplankton feeding pelagic fish (Bulman et al., 2010; Cury et al., 2000;Freon et al., 2009).

There are growing concerns over the ecosystem impacts that areresulting from increased human dependence on fisheries targetinglow-trophic level species (Ainley and Blight, 2009; Branch et al.,2010; Cury et al., 2011; Smith et al., 2011). Most research on theimpacts of these fisheries has focused on high trophic level species,which are increasingly being seen to provide robust ecological per-formance indicators of ecosystem system health, as they are oftensensitive to subtle changes in the abundance and distribution ofthe lower trophic level species on which they prey (see Boyd et al.,2006 for review). Other studies have shown how these fisheriesalso impact on other parts of the marine ecosystem, including othercommercially targeted species (Smith et al., 2011). The growth ofthe sardine fishery since 1991 to become Australia’s largest fisheryhas raised concerns over its potential ecological impacts on otherparts of the marine ecosystem, and the need to develop an EBFMapproach. The current harvest strategy in the fishery is to maintaina baseline total allowable commercial catch (TACC) of ∼30,000 t,while estimates of the spawning biomass fall between 150,000 and300,000 t, corresponding to an exploitation rate of between 20%and 10%, respectively (Ward et al., 2009).

A recent study examined the impacts of low trophic level(forage fish) depletion on global seabird populations, and identi-fied a threshold prey abundance below which breeding successconsistently declined (Cury et al., 2011). This threshold was approx-imately one-third of the maximum observed prey abundance, withthe authors suggesting that a general rule of “one-third for thebirds” could be applied as a guiding principal to EBFM of low-trophic level fisheries, by ensuring that exploitation rates maintaintarget species above one-third of their maximum observed long-term biomass (Cury et al., 2011). Similarly, Smith et al. (2011)showed that broader ecosystem impacts from fishing low trophiclevel species could be substantially reduced by halving exploita-tion rates from typical (∼60%) maximum sustainable yield levels to∼30% exploitation rates. Based on the findings of both these stud-ies, the current exploitation rates in the sardine fishery (10–20%of estimated spawning biomass and ∼24% of the total sardine con-sumption biomass estimated by the EGAB ecosystem model) wouldappear conservative, and results from trophodynamic modellingwould suggest they presently do not significantly impact ecosys-tem function, or high trophic level species. However, it is importantto highlight that most of sardine fishery catch comes from a rela-tively small (∼10,000 km2) area centred on southern Spencer Gulf.This area is ∼20% of the estimated spawning area of sardines inthe EGAB (∼54,000 km2, Ward et al., 2009), which makes up about7% of the total modelled area. The current estimated exploitationrates of sardines in the sardine fishery assumes a high degree ofconnectivity between the sardine population in the main fisheryarea, and other parts of the estimated spawning area within theEGAB. There currently is no data on the movements of sardinesfrom the broader EGAB into the fishery to substantiate this assump-tion. As such, it is possible that exploitation rates are higher withinthe region of the fishery and that localised depletion cannot bediscounted as a meso-scale issue. The model developed and pre-sented in this study covered an extensive geographic area withdiverse habitats and may not be appropriate to examine regionalimpacts from fisheries, such as localised depletion. This will requirethe development of spatially explicit regional ecosystem models,

preferably in conjunction with appropriate ecological performanceindicators (reproductive and foraging success) from land-breedingmarine predators (e.g. seabirds and seals that forage in proximity tothe fishery), to assess and identify issues of localised depletion, and
Page 19: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

5 ogical

efi

itp(napfactvdqppfiarew

A

(2mSasUstwJdaWfGatWAL

A

i2

R

A

A

A

A

6 S.D. Goldsworthy et al. / Ecol

nsure continuing progress is made towards EBFM of the sardineshery.

Results from this study of the EGAB ecosystem highlight themportance of small pelagic fish to the higher trophic levels, therophic changes that result from loss and recovery of apex predatoropulations, and the potential pivotal role of changing cephalopodsquid) biomass in regulating trophic flows. The use of trophody-amic models as a tool to provide context to the potential impactsnd management strategies of a single fishery relative to the tem-oral changes of a complex ecosystem subject to dynamic impactsrom multiple fishing fleets, recovery of apex predator populationsnd climate change is highlighted by this study. Although there arelearly some limitations of the model due to data deficiencies, andhe number and composition of functional groups, the model pro-ides a basis from which future improvements in model design andata inputs will enable more complex management and ecologicaluestions to be examined. The ecosystem performance indicatorsroduced by the Ecosim model also provide a means to assess theotential impacts of the sardine fishery relative to those from othersheries and environmental change. This ability to resolve andttribute potential impacts from multiple fishing fleets and envi-onmental changes is also critical for the development and utility ofcological performance indicators for assessing EBFM targets, andill not be possible without further modelling.

cknowledgements

We thank the Fisheries Research and Development CorporationFRDC) for support of this project (FRDC projects 2003/072, and005/031), and FRDC staff for their support with project manage-ent, especially Crispin Ashby and Caroline Stewardson. We thank

teve Shanks, Will Zacharin, Martin Smallridge (PIRSA Fisheries),nd Christian Pike project support. We thank Topa Petit (Univer-ity of South Australia), Jeremy Austin, Sean Connell (Adelaideniversity), Mike Steer (SARDI Aquatic Sciences) who provided co-

upervisory support to students who assisted in this project. Wehank Alex Ivey and Richard Saunders (SARDI Aquatic Sciences)ho provided additional project support, and Charles James and

ohn Middleton (SARDI Aquatic Sciences) for providing wind stressata. Tom Okey provided help and direction during the conceptu-lisation of the ecosystem modelling components of the project.e thank Peter Gill and Margie Morrice (Blue Whale Study Inc.)

or providing diet data for blue whales and Cath Kemper and Sueibbs (South Australian Museum) provided diet data for bottlenosend common dolphins. We thank Peter Shaughnessy (South Aus-ralian Museum) for provided data on NZ fur seal pup production.

e thank the following for editorial assistance: Peter Shaughnesy,lex Ivey, Cameron Dixon, Shane Roberts, David Currie, Mayleenoo, Adrian Linnane, Jason Tanner and Tony Fowler.

ppendix A. Supplementary data

Supplementary data associated with this article can be found,n the online version, at http://dx.doi.org/10.1016/j.ecolmodel.013.01.006.

eferences

brams, P.A., Ginzburg, L.R., 2000. The nature of predation: prey dependent, ratiodependent or neither. Trends in Ecology and Evolution 15, 337–341.

inley, D.G., Blight, L.K., 2009. Ecological repercussions of historical fish extractionfrom the Southern Ocean. Fish and Fisheries 10, 13–38.

insworth, C.H., Pitcher, T.J., 2006. Modifying Kempton’s species diversity index foruse with ecosystem simulation models. Ecological Indicators 6, 623–630.

lder, J., Campbell, B., Karpouzi, V., Kaschner, K., Pauly, D., 2008. Forage fish:from ecosystems to markets. Annual Review of Environment and Resources 33,153–166.

Modelling 255 (2013) 38– 57

Boyd, I.L., Wanless, S., Camphuysen, C.J., 2006. Top predators in marine ecosystems:their role in monitoring and management. In: Conservation Biology. CambridgeUniversity Press, Cambridge.

Branch, T.A., Stafford, K.M., Palacios, D.M., Allison, C., Bannister, J.L., Burton, C.L.K.,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.,Nishiwaki, S., Sturrock, V.J., Tormosov, D., Anderson, R.C., Baker, A.N., Best, P.B.,Borsa, P., Brownell, R.L., Childerhouse, S., Findlay, K.P., Gerrodette, T., Ilangakoon,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-beek, K., Warneke, R.M., Dolphin Res, G., 2007. Past and present distribution,densities and movements of blue whales Balaenoptera musculus in the SouthernHemisphere and northern Indian Ocean. Mammal Review 37, 116–175.

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. Nature468, 431–435.

Brown, C.J., Fulton, E.A., Hobday, A.J., Matear, R.J., Possingham, H.P., Bulman, C., Chris-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 ofclimate-driven primary production change on marine food webs: implicationsfor fisheries and conservation. Global Change Biology 16, 1194–1212.

Bulman, C.M., Condie, S.A., Neira, F.J., Goldsworthy, S.D., Fulton, E.A., 2010. TheTrophodynamics of Small Pelagic Fishes in the Southern Australian Ecosystemsand the Implications for Ecosystem Modelling of Southern Temperate Fisheries.Final Report to the Fisheries Research and Development Corporation (FRDC2008/023). CSIRO Marine and Atmospheric Research, Hobart, Tasmania.

Chiaradia, A., Costalunga, A., Kerry, K., 2003. The diet of little penguins (Eudyp-tula minor) at Phillip Island, Victoria, in the absence of a major prey – Pilchard(Sardinops sagax). Emu 103, 43–48.

Christensen, V., 2000. Indicators for marine ecosystems affected by fisheries. Marineand Freshwater Research 51, 447–450.

Christensen, V., Pauly, D., 1992. Ecopath II – a software for balancing steady-stateecosystem models and calculating network characteristics. Ecological Modelling61, 169–185.

Christensen, V., Walters, C.J., 2004. Ecopath with Ecosim: methods, capabilities andlimitations. Ecological Modelling 172, 109–139.

Christensen, V., Walters, C.J., Pauly, D., Forrest, R., 2008. Ecopath with Ecosim version6 User Guide, November 2008. Lenfest Ocean Futures Project 2008, p. 154.

Coll, M., Santojanni, A., Palomera, I., Arneri, E., 2009. Food-web changes in the Adri-atic Sea over the last three decades. Marine Ecology Progress Series 318, 17–37.

Copley, P.B., 1996. The status of seabirds in South Australia. In: Ross, G.J.B., Weaver,K., Greig, J.C. (Eds.), The Status of Australia’s Sea birds: Proceedings of theNational Seabird Workshop. Biodiversity Group, Environment Australia, Can-berra, pp. 139–180.

Currie, D.R., Dixson, C.D., Roberts, S.D., Hooper, G.E., Sorokin, S.J., Ward, T.M., 2009.Fishery-independent by-catch survey to inform risk assessment of the SpencerGulf Prawn Trawl Fishery. Report to PIRSA Fisheries. South Australian Researchand Development Institute (Aquatic Sciences), Adelaide. SARDI Publication No.F2009/000369-1. SARDI Research Report Series No. 390, p. 121.

Currie, D.R., Sorokin, S.J., 2010. A preliminary evaluation of the distribution andtrophodynamics of demersal fish from Spencer Gulf. Report to the SouthAustralian Department for the Environment and Heritage. South AustralianResearch and Development Institute (Aquatic Sciences), Adelaide. SARDI Publi-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,H.M., 2000. Small pelagics in upwelling systems: patterns of interaction andstructural changes in “wasp-waist” ecosystems. ICES Journal of Marine Science57, 603–618.

Cury, P.M., Boyd, I.L., Bonhommeau, S., Anker-Nilssen, T., Crawford, R.J.M., Furness,R.W., Mills, J.A., Murphy, E.J., Österblom, H., Paleczny, M., Piatt, J.F., Roux, J.-P., Shannon, L., Sydeman, W.J., 2011. Global seabird response to forage fishdepletion—one-third for the birds. Science 334, 1703–1706.

Dann, P., Norman, F.I., Cullen, J.M., Neira, F.J., Chiaradia, A., 2000. Mortality andbreeding failure of little penguins, Eudyptula minor, in Victoria, 1995–96, follow-ing a widespread mortality of pilchard, Sardinops sagax. Marine and FreshwaterResearch 51, 355–362.

Fletcher, W.J., Chesson, J., Sainsbury, K.J., Hundloe, T., Smith, A.D.M., Whitworth, B.,2002. National ESD reporting Framework for Australian Fisheries: The “How toGuide for Wild Capture Fisheries” FRDC report 2000/145, Canberra, Australia.

Fowler, A.J., Lloyd, M.T., Schmarr, D., 2009. A Preliminary Consideration of By-catchin the Marine Scalefish Fishery of South Australia, South Australian Research andDevelopment Institute (Aquatic Sciences), Adelaide, F2009/000097-1. SARDIResearch 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 relatedto 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,J., Hutchings, L., Masski, H., Mullon, C., Ramdani, M., Seret, B., Simieri, M., 2009.Functional group biodiversity in Eastern Boundary Upwelling Ecosystems ques-tions the wasp-waist trophic structure. Progress in Oceanography 83, 97–106.

Fulton, E.A., 2010. Approaches to end-to-end ecosystem models. Journal of MarineSystems 81, 171–183.

Goldsworthy, S.D., Bulman, C., He, X., Larcombe, J., Littnan, C., 2003. Trophic inter-actions between marine mammals and Australian fisheries: an ecosystemapproach. In: Gales, N., Hindell, M., Kirkwood, R. (Eds.), Marine Mammals and

Page 20: Trophodynamics of the eastern Great Australian Bight ...S.D. Goldsworthy et al. / Ecological Modelling 255 (2013) 38–57 39 (EBFM) objectives (Pikitch et al., 2004). The latter is

ogical

G

G

G

G

G

H

J

J

K

K

L

L

M

M

M

M

M

M

M

M

M

M

M

N

P

S.D. Goldsworthy et al. / Ecol

Humans: Fisheries, Tourism and Management. CSIRO Publishing, Melbourne,pp. 69–99.

oldsworthy, S.D., McKenzie, J., Shaughnessy, P.D., Macintosh, R.R., Page, B., Camp-bell, R., 2009. An Update of the Report: Understanding the Impediments tothe Growth of Australian Sea Lion Populations. Report to the Department ofthe Environment, Water, Heritage and the Arts. SARDI Publication NumberF2008/00847-1, SARDI Research Report series No. 356, p. 175.

oldsworthy, S.D., Page, B., 2007. A risk-assessment approach to evaluating the sig-nificance of seal bycatch in two Australian fisheries. Biological Conservation 139,269–285.

oldsworthy, S.D., Page, B., Shaughnessy, P.D., Linnane, A., 2010. Mitigating SealInteractions in the SRLF and the Gillnet Sector SESSF in South Australia. Reportto the Fisheries Research and Development Institute. South Australian Researchand Development Institute (Aquatic Sciences), Adelaide. SARDI Publication No.F2009/000613-1. SARDI Research Report Series No. 405, p. 213.

riffiths, S.P., Young, J.W., Lansdell, M.J., Campbell, R.A., Hamption, J., Hoyle, S.D.,Langley, A., Bromhead, D., Hinton, M.G., 2010. Ecological effects of longline fish-ing and climate change on the pelagic ecosystem off eastern Australia. Reviewsin Fish Biology and Fisheries 20, 239–272.

uénette, S., Christensen, V., Pauly, D., 2008. Trophic modelling of the Peruvianupwelling ecosystem: towards reconciliation of multiple datasets. Progress inOceanography 79, 326–335.

amer, D.J., Ward, T.M., McGarvey, R., 2008. Measurement, management and miti-gation of operational interactions between the South Australian Sardine Fisheryand short-beaked common dolphins (Delphinus delphis). Biological Conservation141, 2865–2878.

ahncke, J., Checkley, D.M., Hunt, G.L., Harding, J.S., 2004. Trends in carbon flux toseabirds in the Peruvian upwelling system: effects of wind and fisheries onpopulation regulation. Fisheries Oceanography 13, 208–223.

ennings, G., McGlashan, D.J., Furness, R.W., 2012. Responses to changes in spratabundance of common tern breeding numbers at 12 colonies in the Firth ofForth, east Scotland. ICES Journal of Marine Science 69, 572–577.

empton, R.A., Taylor, L.R., 1976. Models and statistics for species diversity. Nature262, 818–820.

night, M.A., Tsolos, A., 2010. South Australian Wild Fisheries Information andStatistics Report 2008/09. South Australian Research and Development Insti-tute (Aquatic Sciences), Adelaide. SARDI Publication No. F2008/000804-2, SARDIResearch Report Series No. 417, p. 64.

ing, J.K., 1999. Exploitation of fur seals and sea lions from Australian, New Zealandand adjacent subantarctic islands during the eighteenth, nineteenth and twen-tieth centuries. Australian Zoologist 31, 323–350.

ozano-Montes, H.M., Loneragan, N.R., Babcock, R.C., Jackson, K., 2011. Using trophicflows and ecosystem structure to model the effects of fishing in the Jurien BayMarine Park, temperate Western Australia. Marine and Freshwater Research 62,421–431.

ann, K.H., Lazier, J.R.N., 1996. Dynamics of Marine Ecosystems. Blackwell Scientific,Oxford, p. 394.

cClatchie, S., Middleton, J.F., Ward, T.M., 2006. Water mass analysis and alongshorevariation in upwelling strength in the eastern Great Australian Bight. Journal ofGeophysical Research 111, 1–13.

cLeay, L.J., Page, B., Goldsworthy, S.D., Paton, D.C., Teixeira, C., Burch, P., Ward, T.M.,2010. Foraging behaviour and habitat use of a short-ranging seabird, the crestedtern. Marine Ecology Progress Series 411, 271–283.

cLeay, L.J., Page, B., Goldsworthy, S.D., Ward, T.M., Paton, D.C., 2009a. Size mat-ters: variation in the diet of chick and adult crested terns. Marine Biology 156,1765–1780.

cLeay, L.J., Page, B., Goldsworthy, S.D., Ward, T.M., Paton, D.C., Waterman, M., Mur-ray, M.D., 2009b. Demographic and morphological responses to prey depletion ina crested tern (Sterna bergii) population: can fish mortality events highlight per-formance indicators for fisheries management? ICES Journal of Marine Science66, 237–247.

erino, G., Barange, M., Mullon, C.n., 2010. Climate variability and change scenariosfor a marine commodity: modelling small pelagic fish, fisheries and fishmeal ina globalized market. Journal of Marine Systems 81, 196–205.

iddleton, J.F., Arthur, C., vanRuth, P., Ward, T.M., McClean, J.L., Maltrud, M.E., Gill, P.,Levings, A., Middleton, S., 2007. El Nino effects and upwelling off South Australia.Journal of Physical Oceanography 37, 2458–2477.

iddleton, J.F., Bye, J.A.T., 2007. A review of the shelf-slope circulation alongAustralia’s southern shelves: Cape Leeuwin to Portland. Progress in Oceanog-raphy 75, 1–41.

iddleton, J.F., Cirano, M., 2002. A boundary current along Australia’s southernshelves: the Flinders Current. Journal of Geophysical Research 107.

iddleton, J.F., Platov, G., 2003. The mean summertime circulation along Australia’ssouthern shelves: a numerical study. Journal of Physical Oceanography 33,2270–2287.

yers, R.A., Baum, J., Sheperd, T., Powers, S.P., Peterson, C.H., 2007. CascadingEffects of the Loss of Apex Predatory Sharks from a Coastal Ocean. Science 315,1846–1850.

aylor, R.L., Hardy, R.W., Bureau, D.P., Chiu, A., Elliott, M., Farrell, A.P., Forster, I.,

Gatlin, D.M., Goldburg, R.J., Hua, K., Nichols, P.D., 2009. Feeding aquaculture inan era of finite resources. Proceedings of the National Academy of Sciences ofthe United States of America 106, 15103–15110.

age, B., Goldsworthy, S.D., McLeay, L., Wiebkin, A., Peters, K., Einoder, L., Rogers, P.,Braley, M., Gibbs, S., McKenzie, J., Huveneers, C., Caines, R., Daly, K., Harrison, S.,

Modelling 255 (2013) 38– 57 57

Baylis, A., Morrice, M., Gill, P., McIntosh, R., Bool, N., Ward, T., 2011. The diets ofmarine predators in southern Australia: assessing the need for ecosystem-basedmanagement of the South Australian sardine fishery. In: Goldsworthy, S.D., Page,B., Rogers, P., Ward, T.M. (Eds.), Establishing Ecosystem-based Managementfor the South Australian Sardine Fishery: Developing Ecological PerformanceIndicators and Reference Points to Assess the Need for Ecological Allocations.Final Report to the Fisheries Research and Development Corporation, PN2005/031.SARDI Publication No. F2010/000863-1. SARDI Research Report SeriesNo. 529. South Australian Research and Development Institute, Adelaide, pp.31–60.

Pauly, D., Christensen, V., Dalsgaard, J., Froese, R., Torres, F.J., 1998. Fishing downmarine food webs. Science 279, 860–863.

Pauly, D., Palomares, M.L., 2005. Fishing down marine food webs: it is far morepervasive than we thought. Bulletin of Marine Science 76, 197–211.

Pikitch, E.K., Santora, C., Babcock, E.A.A.B., Bonfil, R., Conover, D.O., Dayton, P.,Doukakis, P., Fluharty, D., Heneman, B., Houde, E.D., Link, J., Livingston, P.A., Man-gel, M., McAllister, M.K., Pope, J., Sainsbury, K.J., 2004. Ecosystem-based fisherymanagement. Science 305, 346–347.

Piroddi, C., Bearzi, G., Christensen, V., 2010. Effects of local fisheries and ocean pro-ductivity on the northeastern Ionian Sea ecosystem. Ecological Modelling 221,1526–1544.

Polovina, J.J., 1984. Model of a coral reef ecosystem. The ECOPATH model and itsapplication to French Frigate Shoals. Coral Reefs 3, 1–11.

Roberts, S.D., Steer, M.A., 2010. By-product assessment in the Spencer Gulf PrawnFishery with an emphasis on developing management options for Balmain bugs.South Australian Research and Development Institute (Aquatic Sciences), Ade-laide. SARDI Publication No. F2010/000165-1. SARDI Research Report Series No.439, p. 48.

Samhouri, J.F., Levin, P.S., Harvey, C.J., 2009. Quantitative evaluation of marineecosystem indicator performance using food web models. Ecosystems 12,1283–1298.

Shannon, L.J., Neira, S., Taylor, M., 2008. Comparing internal and external driversin the southern Benguela and the southern and northern Humboldt upwellingecosystems. African Journal of Marine Science 30, 63–84.

Shaughnessy, P.D., McKenzie, J., Lancaster, M.L., Goldsworthy, S.D., Dennis, T.E.,2010. Australian fur seals establish haulout sites and a breeding colony in SouthAustralia. Australian Journal of Zoology 58, 94–103.

Shaughnessy, P.D., McKeown, A., 2002. Trends in abundance of New Zealand furseals, Arctocephalus forsteri, at the Neptune Islands, South Australia. WildlifeResearch 29, 363–370.

Smith, A.D.M., Brown, C.J., Bulman, C.M., Fulton, E.A., Johnson, P., Kaplan, I.C., Lozano-Montes, H., Mackinson, S., Marzloff, M., Shannon, L.J., Shin, Y.-J., Tam, J., 2011.Impacts of fishing low-trophic level species on marine ecosystems. Science 333,1147–1150.

Ulanowicz, R.E., Puccia, C.J., 1990. Mixed trophic impacts in ecosystems. Coenoses5, 7–16.

van Ruth, P.D., Ganf, G.G., Ward, T.M., 2010a. Hot-spots of primary productivity: analternative interpretation to conventional upwelling models. Estuarine, Coastaland Shelf Science 90, 142–158.

van Ruth, P.D., Ganf, G.G., Ward, T.M., 2010b. The influence of mixing on primaryproductivity: a unique application of classical critical depth theory. Progress inOceanography 85, 224–235.

Walters, C.J., Christensen, V., Pauly, D., 1997. Structuring dynamic models ofexploited ecosystems from trophic mass-balance assessments. Reviews in FishBiology and Fisheries 7, 139–172.

Ward, T.M., Hoedt, F., McLeay, L., Dimmlich, W.F., Kinloch, M., Jackson, G., McGarvey,R., Rogers, P.J., Jones, K., 2001. Effects of the 1995 and 1998 mass mortality eventson the spawning biomass of sardine. Sardinops sagax, in South Australian waters.ICES Journal of Marine Science 58, 865–875.

Ward, T.M., Ivey, A.R., Mcleay, L.J., Burch, P., 2009. Spawning biomass of sardine,Sardinops sagax, in waters off South Australia in 2009. South Australian Researchand Development Institute (Aquatic Sciences), Adelaide. Final Report to PIRSAFisheries, SARDI Publication No. F2007/000566-3.

Ward, T.M., Mcleay, L.J., Dimmlich, W.F., Rogers, P.J., McClatchie, S., Matthews,R., Kampf, J., van Ruth, P.D., 2006. Pelagic ecology of a northern boundarycurrent system: effects of upwelling on the production and distribution of sar-dine (Sardinops sagax), anchovy (Engraulis australis) and Southern Bluefin Tuna(Thunnus maccoyii) in the Great Australian Bight. Fisheries Oceanography 15,191–207.

Ward, T.M., Mcleay, L.J., Rogers, P.J., 2005. Spawning biomass of sardine (pilchard,Sardinops sagax) in South Australian waters. South Australian Research andDevelopment Institute (Aquatic Sciences), Adelaide. Final Report prepared forPIRSA Fisheries, RD05/0019-1.

Ward, T.M., Rogers, P.J., McLeay, L.J., 2007. Developing and evaluating the use of theDaily Egg Production Method to estimate the spawning biomass of blue mack-erel Scomber australasicus off southern and eastern Australia. In: Ward, T.M.,Rogers, P.J. (Eds.), Development and Evaluation of Egg-based Stock AssessmentMethods for Blue Mackerel Scomber australasicus in Southern Australia. Reportto the Fisheries Research and Development Corporation, Project No. 2002/061,

pp. 311–352.

Wilson, D., Curtotti, R., Begg, G., Phillips, K., 2009. Fishery Status Reports 2008: Statusof Fish Stocks and Fisheries Managed by the Australian Government. Bureau ofRural Sciences and Australian Bureau of Agriculture and Resource Economics,Canberra.