Academia.eduAcademia.edu
REVIEW OF THE MARINE ENVIRONMENT OF THE ALINYTJARA WILURARA NRM REGION, AND SUMMARY OF THREATS (c) A. Loisier REPORT FOR THE ALINYTJARA WILURARA NATURAL RESOURCES MANAGEMENT BOARD by J.L. Baker, Marine Ecologist June 2010 Contents Part 1 AREA DESCRIPTION - MARINE ENVIRONMENT OF THE AW NRM REGION ..................................... 4 1 Area Description .................................................................................................................................... 4 Regional Setting and Biogeography...................................................................................................... 4 Oceanography ...................................................................................................................................... 6 Productivity and Food Webs ................................................................................................................ 8 Geomorphology, Geology and Sediments .......................................................................................... 10 Major Coastal and Marine Habitats.................................................................................................... 14 Benthic Flora ....................................................................................................................................... 18 Invertebrates ...................................................................................................................................... 20 Bony Fishes ......................................................................................................................................... 25 Cartilaginous Fishes ............................................................................................................................ 26 Marine Mammals ............................................................................................................................... 31 Marine Reptiles .................................................................................................................................. 34 Coastal Birds and Sea Birds................................................................................................................. 34 Part 2 USES OF THE AW NRM REGION – MARINE COMPONENT ............................................................. 38 2. Uses and Activities .............................................................................................................................. 38 Recreational Fishing ........................................................................................................................... 38 Commercial Fishing ............................................................................................................................ 42 Coastal and Marine Recreation and Tourism ..................................................................................... 48 Aboriginal Heritage ............................................................................................................................. 49 Research ............................................................................................................................................. 50 Ports, Shipping and Boating ............................................................................................................... 52 Mining, Minerals, Oil and Gas Exploration ......................................................................................... 53 2 Part 3 POTENTIAL THREATS AND THREATENING PROCESSES IN THE AW NRM REGION – MARINE COMPONENT ............................................................................................................................. 57 3 Potential Threats ................................................................................................................................ 57 Fishing ................................................................................................................................................. 57 Threats to Sea Lions............................................................................................................................ 66 Recreation and Tourism ..................................................................................................................... 67 Marine Debris ..................................................................................................................................... 69 Marine Pests ....................................................................................................................................... 70 Research ............................................................................................................................................. 87 Land-based and Coastal Discharges ................................................................................................... 87 Mining, Minerals, Oil and Gas Exploration ......................................................................................... 88 Climate Change ................................................................................................................................... 90 Ranking of Threats and Proposed Actions .......................................................................................... 95 Part 4 CONSERVATION AND MANAGEMENT IN THE AW NRM REGION – MARINE COMPONENT .......... 103 4.1 Notes on Current Protection and Management ............................................................................ 103 4.2 Notes on Proposed Protection and Management.......................................................................... 108 5. Acknowledgements .............................................................................................................................. 110 6. Bibliography .......................................................................................................................................... 110 7. Appendices ........................................................................................................................................... 138 3 Part 1 AREA DESCRIPTION: MARINE ENVIRONMENT OF THE AW NRM REGION 1 Area Description Regional Setting and Biogeography The Alinytjara Wilurara (AW) Region in the far north-west of South Australia, covers 26% of the State, but only a small portion of the region is marine (Figure 1). The marine component stretches from South Australia‘s border with Western Australia in the west, to the coastal edge (Yalata Beach) of the Yalata Indigenous Protected Area in the east, and the regional boundary extends to the edge of the State Waters, 3 nautical miles from the coast (Natural Resources Management Act, 2004). The marine component of the AW NRM region is part of the Statemanaged waters of the Great Australian Bight Marine Park (Figure 1). The AW Region is part of the Great Australian Bight, Australian‘s largest coastal indentation, which extends from Cape Pasley in the west, to Cape Carnot near Port Lincoln in the east, a distance of approximately 1,160kms (DEWR, 2006). The shallow continental shelf in the Bight is very wide. In some places the shelf break, where the water is around 200 metres deep, is over 100 nautical miles (or around 190 kilometres) away from the coastline. Figure 1: Map of the coastal and marine component of the Alinytjara Wilurara NRM Region, showing the boundary of the Region, and the location of the GAB Marine Park. Map provided by AW NRM Board. Major features of the marine area, where the Nullarbor Plain meets the Southern Ocean, are the spectacular Bunda Cliffs (Nullarbor cliffs) of the Great Australian Bight, and the massive sand dunes and wild windswept beaches of the Yalata Lands. The Great Australian Bight is an unusual environment because the coastline is aligned with the equator, and runs mostly from east to west instead of from north to south. The GAB is the longest ice-free coastline running in this direction in the Southern Hemisphere. 4 The South Australian section of the Great Australian Bight (GAB), in which the AW NRM marine region is situated, also forms part of a larger State marine bioregion which extends seaward to the 200nm mark. This area, the Eucla bioregion (Figure 2, IMCRA Technical Group, 1998), is the most westerly of South Australia‘s 8 marine bioregions, and extends westwards past the Western Australian border, to Israelite Bay. The eastern boundary of the Eucla bioregion is situated at Cape Adieu. At an even larger scale, the head of the GAB forms part of the ―Great Australian Bight IMCRA Transition‖ (Geoscience Australia, CSIRO and Department of the Environment and Heritage, undated), a provincial bioregion that extends to south-western Australia. The Eucla bioregion is characterised by moderate to high wave energy, high swells, seasonal warm water currents, a south and south-west facing coastline, and a broad, shallow continental shelf (IMCRA Technical Group, 1998). Seasonal warmer water currents characterise the Great Australian Bight area, and these are considered to influence the type of marine plants, sessile invertebrates, fish species and migratory taxa, that occur in the region (Maxwell and Cresswell, 1981; Edyvane and Baker, 1999; DEWR, 2006). The marine flora and fauna is transitional warm to cool temperate (Flindersian biogeographic province: Womersley and Edmunds, 1958), and there is a distinct tropical and subtropical element to the flora and fauna of the region (e.g. plankton, fish, echinoderms, hydroids) (Markina and Vishnyakova, 1977; Maxwell and Cresswell, 1981, cited by Edyvane and Baker, 1999) due to both locally generated currents, and the edge of the Leeuwin Current from Western Australia. A number of large, sub-tropical pelagic fishes, and also marine turtles, seasonally occur along the coast of the northern GAB, in association with the warm water currents. Various surveys, some results of which are discussed in later sections of this document, indicate that the Head of the GAB supports a diverse mix of habitats, including nearshore reefs dominated by large brown macroalgae; seagrass beds (e.g. near Eucla, west of AW NRM region), and in some of the bays to the east; sponge beds in some of the deeper waters; and mixed reefs dominated by various attached invertebrates, including bryozoans, sponges and ascidians (SARDI data, 1995; Edyvane and Baker, 1998, 1999; Ward et al., 2006; Gurgel, 2009). Furthermore, collation of museum records and previous survey records, indicate that the State-managed waters of the GAB are species-rich and diverse, with pertinent examples including the large numbers of species of bony fishes (at least 332 species), decapod crustaceans (at least 145 species), shelled gastropod molluscs (at least 485 species), echinoderms (at least 123 species), ascidians, bryozoans and sponges (see section 1.7 Invertebrates, and Appendices). Various other groups (e.g. amphipod and isopod crustaceans, amongst many others) are also likely to be species-rich in the GAB shelf waters, but only the few of the many groups of marine invertebrates have been catalogued for this report. 5 Figure 2: Marine Bioregions of South Australia, showing location of Eucla Bioregion in the northern Great Australian Bight (adapted from IMCRA Technical Group, 2008). Oceanography The GAB region is one of high energy, with large wind-driven waves, and strong, year-round south-westerly swells (sometimes to 4m or more) around some of the exposed headlands, and against the cliffs. In winter, this area is fully exposed to storms. In areas away from wave action, the clear waters allow light to reach greater depths than in many other parts of the South Australian coast. A complex mix of weak down-welling currents are layered through the GAB, from the inner shelf, to the deeper waters the slope. These include (i) seasonally alternating flows of small coastal currents, close to the surface; (ii) the eastward flowing warm Leeuwin Current (see below), which joins with the South Australian Current along the southern coast of Australia, and (iii) in deeper continental slope waters, the westward flowing Flinders Current, which brings cooler waters and nutrients into the southern part of the Bight (particularly the shelf break), and has little influence at the Head of the Bight. In addition to the main currents, large eddies and gyres are common in the GAB, whereby cold, oceanic surface water mixes with the warmer waters in the inner Bight. In the summer, the hot, dry winds blow from the deserts of central Australia over the Bunda Cliffs, which form the northern and north-western coastline of the Bight. The winds help evaporate more water from the sea surface, increasing the amount of salt in the surface waters. Evaporation during summer over the shallow expanses at the Head of the Bight, creates hypersaline seawater which generates an additional and predictable down-welling force when the system cools during winter (DEWR, 2006). 6 The coastal area is within the influence of the warm water masses generated at the Head of the Bight, and also by the edge of the Leeuwin Current from Western Australia, which influences the GAB in a weaker way compared with its flow in W.A. waters. The Leeuwin Current is a shallow (< 300m deep) and narrow (< 100km wide) current which originates in the tropical waters of the Indian Ocean, and accelerates as it runs southward down the coast of Western Australia. The Leeuwin Current wraps around the southern corner of the continent and flows eastward along the coasts of Western and South Australia into the Great Australian Bight and further east (DEWR, 2006). This current is considered to suppress upwelling (Hobday et al., 2006) The Leeuwin current is joined by an easterly-moving warm water mass that is generated in the Bight, and the influence of this warm water mass extends east to the Eyre coast (Herzfeld, 1997; Herzfeld and Tomczak, 1997; Herzfeld, 2000). The warm water mass develops in the summer months, resulting in a temperature 2-3°C above that of surrounding waters and spreads in a south-eastward direction to about 137° - 138°E during late summer and early autumn (Herzfeld, 1997). Sampling has shown that during late spring, a large pool of warm water (> 16ºC) characterises much of the inner-shelf south of the Head of the Bight, and is consistent with a feature known as the GAB Plume; the intrusion of heated water that develops due to strong summer heating in the shallow areas of the north-western Bight (Herzfeld, 1997, cited by Currie et al., 2007). Water temperatures gradually decline along the inner-shelf to the east of the warm pool, and reach their lowest values (~15ºC) at the foot of the Eyre Peninsula. This development of warm water at the head of the GAB during summer is considered to be independent of any influence of the Leeuwin Current, and the main source of the warm GAB water is therefore from processes local to the region. The region of greatest warming in the north-western GAB is associated with a large expanse of shallow water (< 30 m), and a larger temperature change in the water column over this region compared with surrounding deeper regions (Herzfeld, 1997). Inshore waters of the middle and mid-eastern part of the Bight stratify in summer, when sea surface temperatures warm to over 23º C, and salinity increases due to evaporation. During late autumn and winter, the inshore waters cool and the Leeuwin Current intrudes into the GAB region from the west, and interacts with the GAB plume over the outer shelf and upper slope (James et al., 2001). A hydrographic survey in summer of 1999 showed that a subsurface front develops over the shelf break during summer, and is indicative of upwelling (Young et al., 1999). This front is capped by warm water. A similar survey by CSIRO also showed the presence of the subsurface front overlain by the cap of warm water, indicating that the vertical structure observed was typical for that time of year. Shelf water is stratified in summer due to localised upwelling and is well-mixed in winter because of swell and storm waves, and the influence of the Leeuwin Current (James et al., 2001). Cooler, high-salinity water (> 36‰) dominates inner-shelf waters, east of the Head of the Bight (Currie et al., 2007). This feature is consistent with evaporative forcing and winter cooling of the eastwardly drifting GAB Plume. Further cooling and winter down-welling of the dense saline water occurs in the eastern part of the GAB, the down-welling appears to be countered in the central GAB by an on-shelf flow of less saline water (< 35.5‰) from the open ocean to the south (Currie et al., 2007). 7 In the GAB, there is also some evidence to suggest that the El Niño - Southern Oscillation (ENSO) climate pattern could also be a key driver that affects productivity and biodiversity in the region (DEWR, 2006). ENSO is composed of (i) an oceanic component, called El Niño (the warm oceanic phase, coupled with La Niña, the cold intervening phase), which refers to a sequence of changes in circulations across the Pacific Ocean and Indonesian archipelago, when oceanic warming is particularly strong (on average every three to eight years) (Bureau of Meteorology, 2010). Characteristic changes in the atmosphere accompany those in the ocean, resulting in altered weather patterns across the globe; and (ii) an atmospheric component, the Southern Oscillation, a major air pressure shift between the Asian and east Pacific regions, whose best-known extremes are El Niño events. Observations suggest that ENSO events significantly influence the strength of currents and upwelling, and the propagation of eddies. ENSO has a strong and very significant effect on the intensity of the southward flowing Leeuwin Current, and Australia‘s west coast waters, and this signal in the Leeuwin Current is further transmitted along the south coast of Australia (Holbrook et al., 2009). The Leeuwin Current is observed to be strongest during the La Niña phase, as are the mixing mechanisms of the winter overturn; westerly wind strength; number and strength of storm fronts; and rates of surface cooling (DEWR, 2006). Productivity and Food Webs The GAB is considered, in oceanographic terms, to be an area of low productivity at a large (continental) scale, due to (i) the absence of high-flowing river systems and the consequent limited amount of terrigenous (i.e. land-based) nutrient inputs; and (ii) the combined absence of large-scale upwelling, and the predominance of seasonal down-welling (probably as a result of the influence of warm water currents – see 1.2 Oceanography) (McClatchie et al., 2006; DEWR, 2006). The interactions between the Leeuwin Current, other currents and the continental shelf have implications for the productivity and ecology of the GAB, and may influence localised upwellings. The ‗conveyor belt‘ system made up by the Leeuwin Current (flowing eastward) and the deeper Flinders Current (flowing westward) is likely to be used for large-scale movements by a number migratory species (DEWR, 2006). Summer upwellings increase the productivity of shelf waters in parts of the GAB (Young et al., 1999), as occurs in other parts of South Australia, but the upwellings in the Head of Bight area are weaker than those observed further south-east. Chlorophyll density, a measure of primary productivity of phytoplankton, is high in summer in the Fowlers Bay area (Ward et al., 2008), just east of the AW NRM region. Correspondingly, during a summer survey in 1999, biomass of zooplankton estimated by bongo net tows was highest inshore in the GAB (compared with offshore) and consisted mainly of a tropical salp, and biomass of the larger micro-nekton estimated by a mid-water trawl net, was highest in offshore waters (Young et al., 1999). Studies of fluorescence (an indirect indicator of the primary productivity), zooplankton abundance, and abundance of eggs of Australian pilchard (Sardinops sagax) and anchovy Engraulis australis across the central and western South Australian coast, have shown that in some years, these measures are seasonally high at Head of the Bight (and in shelf waters to the south, south-west 8 and south-east of that point) and also at many sites in eastern GAB (Ward et al., 2001, 2005; Dimmlich et al., 2004). Schools of pilchards are commonly associated with small benthic ―hills‖ and other geomorphic features in inshore waters. A thermo-cline occurs at approximately 40m across the shelf, and appears to influence species distribution, and potentially plays a role in nutrient upwelling (DEWR, 2006). Also, in some parts of the Head of the Bight, groundwater sources are hypothesised to play some role in augmenting the productivity of coastal ecosystems (DEWR, 2006). Apparently in the Head of the Bight, intrusions of submarine groundwater / seeps may influence the community composition, but the importance of these submarine groundwater seeps needs further investigation. Satellite images show higher concentrations of chlorophyll a in the Head of Bight area compared with waters directly east or west of this area. The localised increase in productivity may be supported by anecdotal observations of higher concentrations of a number of species such as juvenile west Australian salmon, mulloway, school shark, Australian sea lion, and dolphins, all of which use the inshore habitats as nurseries, and, for some of these species, feeding grounds (DEWR, 2006). A study by CSIRO (cited by Young et al., 1999) showed that the summer front in the GAB (see 1.2 Oceanography) is correlated with a relatively higher biomass of potential prey of southern bluefin tuna (Thunnus maccoyii). Research has shown that juvenile southern bluefin tuna predominantly feed on clupeoids, fishes in the sardine and herring family (Young et al., 1997; Ward et al., 2006). Pilchard stocks may thus be highly significant for juvenile southern bluefin tuna in the Great Australian Bight and the southern Indian Ocean. Individual Bluefin Tuna of about one to four years old are found seasonally in the surface waters off southern Australia, and the Bight appears to be an important feeding and nursery ground during the summer (Cowling et al. 1996, cited by Director of National Parks, 2005; Figure 12 in Ward et al., 2006). Possibly due to relatively low productivity, the GAB supports mainly higher order predators that are highly mobile and transitory, rather than resident. Highly mobile predators that are characteristic of marine food webs in the GAB, include marine mammals (southern right whales, Australian sea lions, dolphins, and some toothed whale species), sharks (e.g. white shark, bronze whalers and dusky shark / black whaler, amongst others) and piscivorous fish (e.g. southern bluefin tuna, yellowtail kingfish, west Australian salmon, pink snapper, and deeper water species such as gemfish) and a variety of seabird species (DEWR, 2006). Many of these species are opportunistic predators, that follow frontal features between areas of enhanced productivity, or pursue schools of smaller bait fish (e.g. sardines, anchovy, herring) as they migrate within the Bight. Many of the larger predatory species in the GAB are flexible in their diet and prey selection, perhaps to exploit the seasonal availability of different food sources in the inshore waters, and this is probably linked to variability in the dominant physical features of the GAB such as the warm water currents, and winter storms, and wave energy (see 1.2). Life stage-specific migrations of marine species have important implications for the dynamics of food webs in GAB. Many marine species are observed to move between shallow and deeper water environments during the course of their lifecycle. Typically, such species spend their juvenile stages in shallower water, feeding on and amongst smaller animals like themselves. As sub-adults they may feed across the flats of the shelf, gradually moving offshore towards adult feeding grounds at the shelf edge, on the slope, or in oceanic waters. Examples include western rock lobster, pink snapper, some of the ground sharks, and other shark species (DEWR, 2006). 9 Baitfishes have a significant role in the food web of the GAB, and the size and variability of the biomass of small pelagic fish such as pilchards, may have significant influence on the distribution and biomass of other fauna (Ward et al., 2008). For example, baitfishes are an important food source for young southern bluefin tuna. Shark fishers in the GAB have described pulses of ―dirty water‖ that irregularly wash through the GAB from the west. These pulses are reportedly followed by an increased availability of baitfish and squid and, later, larger predators (e.g. dolphins, tuna, school sharks and whaler sharks) (DEWR, 2006). Many of the seasonal predators, such as tuna, Australian salmon, and little penguin, feed on baitfishes (Fletcher, 1990). During this time, many species tend to concentrate along the shelf edge. These productivity pulses are highly variable within and between years, and may be associated with La Niña years. Seabirds are attracted to the area by the ephemeral blooming of shelf edge plankton communities and they then track the food chains associated with these pulses of productivity as they move through the system. Other species that are common in GAB shelf waters may also be responding to these events (e.g. ocean leatherjackets, latchet, stingrays, stingarees, gummy shark and angel shark). Geomorphology, Geology and Sediments The Head of the GAB, as part of the Eucla Bioregion, is dominated by a Tertiary limestone cliff coastline (with narrow intertidal rock platforms at the base of the cliffs in some places); and also supports Pleistocene dune rock headlands and reefs in the east, interspersed with Holocene beaches and dune barriers (IMCRA Technical Group, 1998). More specifically, the eastern side (region between Fowlers Bay, and Twin Rocks at the Head of the Bight), is defined geomorphologically as a curving, geologically stable and uniform province, dominated by beach and semi-stable to unstable Holocene dune barrier deposits, regularly interrupted by Pleistocene dune calcarenite cliffs, shore platforms and reefs (Short et al., 1986). A major feature of the area is the large sand drifts, in particular the Yalata dunes which extend 13kms inland (Short et al., 1986). A separate geomorphic province has been defined for the area west of Twin Rocks, extending to the W.A. border. This province is dominated by the Nullarbor cliffs, which extend unbroken for 179kms. These cliffs, known as the Bunda Cliffs at the Head of the Bight, are composed of Tertiary limestone averaging 70m in height (and the cliffs are as high as 90m in some places). About 30 km from the S.A./W.A. border, the cliffs are fronted by a beach and dune calcarenite ramp, mantled by extensive drifting sand dunes, known as the Merdayerrah Sandpatch (Short et al., 1986). Much of the coast in the northern and western section of the Head of the Bight is dominated by the Nullarbor Plain, the onshore part of the sedimentary Eucla Basin. The Nullarbor Plain has an area of ~200,000 km2, making it one of the largest continuous karst areas in the world (Lowry and Jennings, 1974, cited by Webb and James, 2006). Karst topography refers to a landscape shaped by the dissolution of a layer or layers of soluble bedrock, usually carbonate rock such as limestone. The Eucla Basin formed part of the Miocene (less than 65 million years ago) sea-floor, which was subsequently uplifted to form the present day plateau. The surface is flat and very gently undulating; any change in the surface relief is due to karst development. Karst landforms are characterised by solution sculpturing of outcropping rock, underground drainage systems and caverns (ANRA, 2009). Rainfall onto the almost flat surface of the Nullarbor Plain seeps through the limestone, influencing the tunnel and cave systems below. 10 The Nullarbor region was inundated by the sea following subsidence during the early Cretaceous, and the first limestones formed about 45 million years ago, in the Middle Eocene (Lowry, 1970, cited by Bradbury and Eberhard, 2000). There were at least two periods of recession during the Oligocene and early Miocene, followed by uplift about 15 million years ago, during the middle Miocene (Lowry and Jennings, 1974, cited by Bradbury and Eberhard, 2000). Since that time, the Nullarbor Plain has not been subject to marine transgression until the Late Pliocene – early Pleistocene, when sea level reached about 30m above its present level. The limestone plateau of the Nullarbor Plain slopes gently south, terminates abruptly as wave-cut limestone cliffs along the Great Australian Bight. The cliffs along the Nullarbor fall sheer into the sea except in two areas in the centre and west, where there are coastal plains (Webb and James, 2006). In the western section of the Nullarbor, west of the AW NRM region, old sea cliffs (eroded during the late Pliocene – early Pleistocene transgression) mark the northern extent of plains which are now some 30km inland of the present coastline (Bradbury and Eberhard, 2000). At least 100 caves have been recorded along the Nullarbor (Webb and James, 2006). In the limestone karst regions below the Nullarbor there are both dry caves and wet caves (the latter of which intersect the regional water table). Some caves near the coast contain seawater which has flowed in through the permeable rock. Weathering has eroded the limestone plain to produce landscape features such as dolines large circular depressions that are found throughout the karst surface. More than 150 collapse dolines are present on the Nullarbor Plain, mostly within 60 km of the coast, as steep-sided, closed depressions often partially or wholly walled by cliffs, which may be overhanging. Many dolines are degraded, but some are still actively collapsing. Caves in the Nullarbor, like the dolines, are mostly restricted to the coastal belt; there are ~100 that have significant passage lengths. The coastal strip of limestone containing most of the caves is about 400 km long and 50 km wide (Webb and James, 2006). Rock holes are another smaller karst feature, where surface water collects within the dry plain. In some of the wet caves, the salinity is the same as sea water, but the water is enriched with calcium. Examples such as Cocklebiddy in the west, and Warbla and Weebubbie along the eastern margin of the Nullarbor, feature lakes of blue water, and huge passages running under the Nullarbor Plain (Yalata Land Management, 2003). Caves within the AW NRM region include Koonalda (north of the Nullarbor Roadhouse), part of the Murrawijinie caves; Warbla; Ivy, and New caves, and Biduna Blowhole (Webb and James, 2006; Atlas of South Australia, 2005). The evolution of the Nullarbor karst system, including the caves, is discussed in detail in Webb and James (2006). Little rain falls on the Nullarbor Plain, and no major rivers flow into the Bight. Therefore, very little sediment washes from the land onto the continental shelf. Seaward of the Nullarbor, there is a shallow offshore gradient in the area, and the continental shelf is dominated by bioclastic carbonate sediments (i.e. formed by skeletal remains of marine organisms) (IMCRA Technical Group, 1998). South from the Head of the Bight, the seafloor is relatively flat, and slopes gently for about 260 km before reaching the shelf edge at 200m depth (Currie et al., 2007). 11 The shelf of the Great Australian Bight lies seaward of the Tertiary Eucla Basin and much of it is covered by coarse calcareous and mainly relic Pleistocene sands that are constantly shifted and rippled by bottom currents. Bryozoans, molluscs, sponges, foraminifera, calcareous red algae, and other calcareous species that live on the shelf comprise the bulk of the relic sands (Conolly and Von Der Borch, 1967). Surface sediments are a mixture of the calcareous Pleistocene skeletal and lithic intraclasts (i.e. relict grains, formed by the re-deposition of material eroded from an original deposit), and Holocene biofragments, with minor amounts of quartz (James et al., 2001). Cool-water carbonate depositional processes have characterised the sea floor, particularly during the Pleistocene, and the GAB forms part of the largest area of cool-water carbonate shelf deposition on the globe, which stretches across the southern continental margin of Australia (Feary and James, 1995). The zones of currently active carbonate production on the shelf occur well east (western Eyre Peninsula) and west (south-western W.A.) of the Head of the Bight (James et al., 2001). According to Feary et al. (2004), the high accumulation rates of sediment, comparable to rates in warm-water carbonate environments, reflect partitioning of sedimentation between the shelf and slope, as high wave energy from the Southern Ocean interacted with sea level fluctuations to generate vigorous off-shelf transport (Feary et al., 2004). The huge middle portion of the GAB shelf is a "shaved shelf" with active sediment winnowing, and mostly relict sediment. In this central GAB region where year-round down-welling occurs, fine sediment is transported off the shelf, and carbonate muds are deposited (James et al., 2001). The outer shelf and upper slope is described as a ―variably productive sediment factory‖ characterised by prolific calcareous growth of benthic organisms such as bryozoans, on hard substrate "islands", which shed particles into surrounding sands and mud (James et al., 2001). Sedimentary facies (i.e. layers distinguished from each other by appearance or composition) in the GAB have been classified by James et al. (2001) (Figure 3 below), who suggested that the sedimentary facies reflect the spatial distribution of benthic assemblages in the GAB. The facies within the marine component of the AW NRM region include the ―Molluscs – Intraclast‖ sand at the Head of the Bight and further offshore, and the ―Quartzose Skeletal‖ sand and gravel facies to the west, extending over the W.A. border. The Molluscs – Intraclast facies comprises wellsorted fine to medium sands and poorly sorted fine sandy gravels, with minor bryozoans and foraminifera. The Quartzose Skeletal facies is characterised by equal amounts of bryozoans and bivalves with lesser quartz, feldspar and crystalline rock fragments. It varies from poorly sorted coarse gravelly sand to rippled fine-to-medium sand (James et al., 2001, cited by Richardson et al., 2005). Bivalves are common in the sediments of the inner shelf, and in shallow shelf regions the sediments comprise bivalve-rich, bioclastic sands (James et al., 1994). Other sediment components include tests of the large benthic foraminifera Marginopora sp. (Feary et al., 1993, cited by Richardson et al., 2005). Water depth plays an important role in determining the textural composition of sediments, and surveys have shown that spatial patterns in grain-size are broadly consistent with patterns in shelf bathymetry (Currie et al., 2007). Sediments are typically coarsest in shallow inshore waters, and become progressively finer with increasing depth and distance offshore (Currie et al., 2007). Recent surveys during the 2000s (Currie et al., 2007) support the previous works (Conolly and Von Der Borch, 1967; James et al., 2001) in 12 confirming that sediments of the GAB shelf are composed almost entirely of biogenic material, such as fragments of bryozoans, molluscs, coralline algae and foraminifera. During the early 2000s, a SARDI Aquatic Sciences project characterised the sea floor of the GAB using videography (Currie et al., 2007). The seafloor of the inner-shelf (40-60 m depth) is reported to be typically composed of hard-packed winnowed sand, swept into irregular, sharpcrested, ripples (wave length = 0-20 cm; amplitude = 0-50 cm). These sands are reportedly mainly bare, but dense patches of epifauna (particularly sponges and ascidians) occur sporadically, stretching for more than 10m. By comparison, high-relief ―sand dunes‖ (wave length = 60-100 cm; amplitude = 20-30 cm) are reported to characterise the bed-forms further offshore (70-120 m depth). These dunes are reportedly reworked by prevailing south-westerly swells, and are characteristically sinuous-crested; with peaks composed of fine sediments and troughs comprising mainly of coarse abraded shell fragments. Epibenthic growth is sparse in this area of the shelf, with small isolated clusters of sponges, ascidians and hydroids occurring almost exclusively in the dune troughs (Currie et al., 2007). On the north-western side of the GAB, in shallow waters, prolific rhodoliths occur, where shallow summer waters are the warmest in the GAB (James et al., 2001). Rhodoliths are unattached calcareous marine plants that look like stones. The warm, saline, nutrient-depleted waters from the north-west then drift eastward across the shelf, suppressing biogenic carbonate production on the central and eastern mid-shelf. This arrested production in the eastern GAB is countered locally by summer coastal upwelling along western Eyre Peninsula, with bryozoanrich sediment extending well inboard onto the mid-shelf (James et al., 2001). Figure 3: Distribution of James et al.’s (2001) surface sediment classifications in the Great Australian Bight (from Richardson et al., 2005). 13 Major Coastal and Marine Habitats The coastal features of the Head of the GAB are discussed in detail in a number of other reports, most recently in the Far West Coastal Action Plan and Conservation Priority Study (Caton et al., 2007), which divided the coast into distinct cells, each of which was photographed and described. Features of the coast are summarised briefly here (Table 1). The marine habitats of the AW NRM region have not been systematically surveyed, due to the relative inaccessibility of the area, as discussed below. However, based on the limited number of nearshore marine surveys to date (see 1.6 Benthic Flora), a brief summary of nearshore marine habitats in the area would include: • rugged, wave-exposed cliff areas (mainly calcareous), including the Bunda Cliffs west of the Head of the Bight, extending for 200km to Wilson Bluff. The cliffs range in height mainly from 40-70 metres above sea level (Yalata Land Management, 2003), up to 90m in some parts; • granitic-based, calcarenite-topped headlands (in the east); • wave-exposed beach habitats, in some areas backed by a large and dynamic dune system, and fronted by sand bars. The winds, waves and currents that feed the dunes vary in strength seasonally, and inter-annually. • wave-exposed bays, containing both seagrass beds and macroalgae-dominated calcareous patch reefs; • nearshore reefs (both granitic and calcareous) of high and medium wave exposure, dominated by canopy brown macroalgae and a variety of large red macroalgae, green macroalgae and turfing brown macroalgae (see descriptions in 1.6 Benthic Flora, below); • island reef habitats to the east (both granitic and calcareous); • offshore calcareous reef and sand habitats, with sponges, bryozoans, molluscs and other benthic fauna. Significant coastal habitats include low cliffs and transgressive dunes from Wahgunyah (east of the AW NRM boundary) to Twin Rocks; the supra-tidal saltmarsh of the Yalata swamp; the large, transgressive sand mass of the Yalata Dunes; the Head of Bight (Bunda Cliffs); the long massive cliff rampart of the Nullarbor, and, in the far west, Merdeyarrah Sandpatch (Yalata Land Management, 2003; Caton et al., 2007). Much of the coastal area west of the Head of the Bight is remarkably uniform, consisting of vertical limestone cliffs. The coastal area of the AW NRM has been mapped in recent years (Caton et al., 2007). A description of coastal environments is summarised in the table below. 14 Table 1: Description of coastal and nearshore areas, in 8 – 10km sections from the eastern to the western coastal edge of the AW NRM region, according to Short et al., 1986 and Caton et al., 2007. Cell No. (section of AW NRM coast) AW7: Yalata Beach Approx. Length (km) ~8.5km AW8: Yalata IPA, between Coontiama Well and Bilbabbi Well. ~11km AW9: Yalata IPA, NW of Coontiama Well. ~11km AW10: Yalata IPA, Tjitji Tjutaku and Geues camp sites. ~10.5km AW11: Yalata IPA, Catacombs Cave and Bob‘s Kitchen and Jaxons camp sites. ~8.5km AW12: Yalata IPA, to Head of the Bight / Eyre Well ~10km Description (from Short et al., 1986 & Caton et al., 2007) Open dissipative sandy beach with multiple inshore bars, with some low calcarenite headlands, small bays and low tide reefs in the SE section, near Yalata Beach. Undulating calcarenite coastal plain is overlain by complex, generally transverse dunes, which appear stable inland; however, immediately behind the beach most dunes are unstable. Dissipative beach, with small patches of calcarenite intertidal platform and sub-tidal reefs. Calcarenite frequently exposed as a shoulder at the back of beach. Extensive calcarenite coastal plain, covered with dune sands, with predominantly transverse forms. Dunes within 500m of the beach are generally bare and de-stabilised. High to medium energy beach along the whole section, in places Pleistocene calcarenite is present in small platforms and nearshore reefs, and in low headlands, or outcropping at the back of the beach. The calcarenite forms a low undulating coastal plain, over which Holocene (white) and Pleistocene (pale brown red) dunes have drifted, and extend across the whole section. The dunes are un-vegetated near the beach: the area of bare sand increases towards the west. Moderate energy beach extends as a series of open embayments along the whole section, separated by low headlands of Pleistocene calcarenite, which also form a low, straight, nearshore reef cut into sections with sand between outcrops of reef. Dunes cover the undulating coastal plain. White, Holocene, dunes are found within ~ 1 km of the shore; these are generally un-vegetated SE of the Tjitji Tjutaku camp site. Relatively straight, medium to high energy, dissipative beach, with 1 to 2 nearshore bars; some low calcarenite headlands and low tide platforms. At the rear of beach, Holocene sand dunes with generally transverse patterns extend inland < 200m. Pleistocene dune calcarenite low cliffs, shore platforms, wave-exposed sand beach and nearshore island (Twin Rocks). Dunes of the vast Yalata sand mass have extendedmore than 7 km inland, as sand has been supplied during Pleistocene and Holocene high sea level episodes to the broad embayment of the Head of Bight. Sand supply has cut off possible former marine connection to the stranded samphire of the low- lying Yalata Swamp. 15 Table 1 (continued): Description of coastal and nearshore areas, in 8 – 10km sections from the eastern to the western coastal edge of the AW NRM region, according to Short et al., 1986 and Caton et al., 2007. AW13: Yalata IPA, western side of Head of the Bight ~9.5km AW14: Yalata IPA, western side of Head of the Bight AW15: Yalata IPA, western side of Head of the Bight, near eastern edge of Nullarbor National Park AW16: eastern end of Nullarbor National Park AW17 – AW19: cliffs adjacent to western side of GAB Marine National Park AW20: cliffs adjacent to western side of GAB Marine National Park AW20 – AW32: cliffs adjacent to western side of GAB Marine National Park AW33: cliffs adjacent to western side of GAB Marine National Park ~9km The coastal land from a little west of Twin Rocks is an almost flat limestone surface covered by skeletal to absent marl and sand soils, broken only by small sink holes. The cliffs are a vertical rampart up to 90m high, extending from the Head of Bight 170km to the Merdayerrah Sandpatch and the WA border. Near the cliff edge there are small degraded cliff top dunes of Pleistocene age, also occurring as small patches throughout the Nullarbor. The limestone of the plateau is Miocene Nullarbor Limestone, a grey- brown hard crystalline rock. This is underlain by the thin chalky Abrakurrie Limestone (Oligocene/ Miocene age); this is underlain in turn by the Eocene Wilson Bluff Limestone: a white friable limestone. The soft unresistant nature of the Abrakurrie and Wilson Bluff Limestones results in marine undercutting and cliff collapse, resulting in vertical or overhanging cliffs. Following a cliff collapse, debris slopes of large limestone blocks clothe the lower cliffs; the frequency of fresh talus/ debris slopes and reports of surface cracking in the Nullarbor Limestone indicate the active erosion and recession of the cliffs. (as for AW13 above) ~9km (as for AW13 above) ~9km (as for AW13 above) ~ 8.5 - 9km (each cell) (as for AW14 above) ~10km (as for AW13 above) ~ 9km 9.5km (as for AW13 above) ~9km AW34: Merdayerrah Sandpatch ~10km The Nullarbor cliffs continues from the east as a degraded cliff line, fronted by a Pleistocene calcarenite slope. The calcarenite slopes are heavily dissected by gullies, and leads down to low cliffs, open embayments, small low tide reefs and small beaches. Merdayerrah Sandpatch consists of Pleistocene dune calcarenite ramps partially covering the backing limestone cliffs, including a 10 km section where Holocene Dunes (the sandpatch) have thinly blanketed the Pleistocene ramps and extend up the limestone cliffs on to the backing plain. The plain surface is relatively flat, at 70 to 90m, with low relict cliff top dunes, thin soils and kunkar (i.e. a nodular calcium carbonate deposit) over the Tertiary limestone cliffs. Extensive shallow sub-tidal reefs are located immediately offshore. 16 It is notable that there are no rivers or estuaries along the central coast of the Great Australian Bight. Some of the marine habitats in the area are discussed briefly below. Data are patchy, because few marine benthic surveys have been undertaken in the shallow shelf waters of the Head of the Bight, due to the remoteness (and therefore the expense) and the relative inaccessibility due to extreme oceanographic conditions for most of the year, which includes high waves and swells, and frequent nearshore rip currents. A partial survey of the marine benthic environment at the Head of the Bight was undertaken in 1994, and several sites were surveyed (Edyvane and Baker, 1998, 1999), as described below in 1.6 Benthic Flora. These included highly wave-exposed reefs at the Head of the Bight, and exposed reefs at Pilpuppie Well and Cape Adieu, the latter east of the AW NRM boundary. Groups of replicate samples at these sites were taken between 10m and 15m deep. Just east of the Head of the Bight, along the Yalata coastline, there is a series of named campsites, and near all of these there is dissected reef forming ―gutters‖ near shore. In order from west to east, these include Hilton, Bob‘s Kitchen, Jaxons, Geues Hole, Tjitji Tjutaku, and Coombra (Coymbra), Granites, and Yalata Beach (Yalata Land Management, 2003). Jaxons is a steeply sloping deep beach west of Geues, and contains nearshore reefs on either side, about 1.5km apart. Geues Hole is a crescent-shaped beach about 200m wide, which is partially sheltered by a series of limestone reefs with a channel (rip) to the sea. The sea flows into Geues Hole through another gap in the reef to the east, forming a distinct channel between the sheltering reef and a small but prominent sand spit, where the water is deeper (Wilson, 2007). There is also reef offshore from Tjitji Tjutaku, a beach about 2km east from Geues. Coombra (= Coymbra) contains two spits of reef nearshore, with an opening to the sea in between. Figure 4 shows the location of beaches (including major camp sites) in the AW NRM region. Figure 4: Location of major camping grounds in AW NRM region, and coastal habitat types. Map provided by AW NRM Board. 17 The coast in the Head of Bight area is dominated by beach and semi-stable to unstable dune barrier deposits, regularly interrupted by dune calcarenite cliffs, shore platforms and reefs. A major feature of the area is the large sand drifts, in particular the Yalata dunes which extend 13kms inland (Short et al., 1986). Further seaward in the region, filter-feeding invertebrate animals, such as sponges, ascidians and bryozoans, are a dominant feature of the sea floor in the shelf waters of the Eucla Bioregion, and these invertebrates, particularly the larger sponges, form 3-dimensional structure on the sea floor. For many fishes and mobile invertebrates in the Great Australian Bight, invertebrate-covered sea floor provides areas for feeding, breeding, camouflage and shelter. Benthic Flora Surveys of the marine benthic flora of the Head of the Bight were undertaken by SARDI Aquatic Sciences at several sites in 1994 (Edyvane and Baker, 1998, 1999), and by the Plant Biodiversity Centre / South Australian Herbarium in 2008 (Gurgel, 2009) and 2010 (F. Gurgel, Plant Biodiversity Centre, pers. comm.). It is important to note that the surveys had different aims, methods, and sampling sites, and therefore the results are not directly comparable. The 1994 survey aimed to characterise subtidal benthic habitats (using SCUBA, at 5m – 20m) over a broad scale of more than 600km (western Eyre Peninsula and the GAB), based on dominant flora that contributed to the percentage cover in replicated 1m square quadrats (Edyvane and Baker 1998). During that survey, 3 sites in the AW NRM were sampled (limestone reef at Head of the Bight; ―Callosity Point‖, and Pilpuppie Well / Dog Fence, ranging from 10m to 15m deep). Several sites east of the AW NRM region, at the Eucla Bioregion boundary, were also sampled (Cape Adieu, Nuyts Reef, Cape Nuyts and three sites in Fowlers Bay, collectively from 5m to 15m deep). The 1994 survey did not aim to sample species richness in the GAB, but it is noted that during the representative sampling to characterise habitats, 40 species were recorded from 20 quadrats at the three sites from 10m to 15m deep within the AW NRM region, considerably fewer than the number of species recorded in a comparable number of quadrats at sites further east of the AW NRM boundary. The 2008 survey aimed to assess the status of the genus Sargassum in the GAB, and to collect as many species as possible from 5 sites in the shallow subtidal (on snorkel), as part of a marine plant species inventory for the AW NRM region (Gurgel, 2009). Therefore, only the second survey (2008) aimed to sample species richness at the Head of the Bight, and due to the major differences cited above, the results cannot meaningfully be compared with the first (1994). The sites sampled during the 2008 survey were off beaches at East Coymbra, West Coymbra, 2 sites at Mexican Hat, and Wandilla Beach, both near Fowlers Bay, east of the AW NRM region boundary. Many of the understorey species recorded during the 1994 survey were not recorded during the 2008 survey, due to differences in survey aim (dominant cover, versus drift and shallow subtidal collection for species richness assessment), methods (SCUBA on highly wave-exposed subtidal reefs versus snorkel from beaches in bays), sampling locations (none were the same between the two periods) and depths (shallow subtidal less than 2m, compared with 10-20m deep). 18 A summary of the dominant cover at subtidal reefs sampled in and around the AW NRM in 1994 is provided below. Additionally, for a summary of the species recorded at 5 shallow subtidal locations off beaches in the AW NRM, readers should refer to Gurgel (2009). Generally, at the Head of the Bight, on high energy limestone reefs, subtidal macroalgae communities are dominated by the brown canopy species Scytothalia dorycarpa and Ecklonia radiata, with species of Cystophora (such as C. platylobium) as lesser cover. Although there are few seagrass communities along this high energy coast, some seagrass beds occur west of the Head of the Bight, in the area around Eucla. Amphibolis antarctica and A. griffithii are both present in that area, with many common seagrass epiphytes, such as Haliptilon roseum and Colpomenia sinuosa (G. Belton, University of Adelaide, pers. comm., 2010). Mixed seagrass and patch reef communities also occur in the shallow subtidal west of the Head of the Bight. Cover at several specific sites is discussed below. Within the marine component of the AW NRM region, benthic samples taken in 1994 at both ―Callosity Point‖ and the Head of the Bight showed that wave-exposed reef in the area was dominated by Ecklonia radiata and Scytothalia dorycarpa, with a sparse understorey of encrusting corallines; the articulated coralline Amphiroa sp., and patches of bare limestone reef and sand. There were fewer species of macroalgae at the very Head of the Bight compared with slightly less wave-exposed areas to the east (discussed below). The existence of strong swell and wave conditions for most of the year, possibly restricts the establishment of many species of macroalgae on the large expanses of bare, mobile sand between reef patches in the inshore area to 20m. As well, the existence of warm water masses at the Head of the Bight may contribute to the low number of species of macroalgae recorded in this region, compared with slightly cooler areas of nutrient-rich upwelling to the east of the Head of the Bight (Edyvane and Baker, 1998). Further south-east, near the eastern boundary of the AW NRM region, benthic samples taken at Pilpuppie Well in 1994 showed that reef in the area was dominated by the canopy species Scytothalia dorycarpa, with less cover of Ecklonia radiata and a Sargassum species (paradoxum?). The understorey was species-rich and diverse, and consisted of large red taxa, such as species of the succulent reds Trigenia and Cladurus; and species of Dictyomenia and Vidalia, which have serrated laterals. Other red species in the understorey suite included species of the ―leafy‖ reds Lenormandia and Crassilingua, the ―pelty‖ Haloplegma and two species of Thuretia, as well as species of Rhodopeltis and Callophyllis. Encrusting corallines were also present (Edyvane and Baker, 1998, 1999). Collectively, the two surveys from 1994 and 2008 have sampled less than 10 sites in the AW NRM region, but further surveys are being undertaken in 2010 (F. Gurgel, pers. comm., 2010). According to a recent analysis of data from the South Australian Herbarium, 200 species of marine plants have been recorded to date in AW NRM region (Table 2; Baker and Gurgel, in prep.), and 31 of these were recorded for the first time in the region during the 2008 survey in the shallow subtidal. The 2008 survey, which included samples taken on snorkel from 5 sites in and adjacent to the AW NRM, recorded 74 species of marine plants, many of which were small epiphytes from shallow waters (Table 3 in Gurgel, 2009). Due to its remoteness and inaccessibility, the AW NRM region has been surveyed far less frequently that other regions of the State, and comparatively little is known of the marine flora compared with well sampled areas. Table 2 compares the number of species of macroalgae recorded to date (2009), and number of records in the South Australian Herbarium database. Nomenclature of all taxa in the 19 SA Herbarium database, including old records, has been corrected and updated where necessary (e.g. Baker and Gurgel, in prep.). the low number of recrods and species recorded in the area reflects the remotenss and inaccessibility of the areas. There is no evidence to indicate that the AW NRM region is one of low diversity or abundance in macroalgae, and recent (e.g. F. Gurgel and G. Belton, unpubl. data) and future targeted collecting efforts may substantially increase the number of species and records known from reefs and seagrass beds in the area. Table 2: Number of species of macroalgae recorded to date (2009), and number of records in the South Australian Herbarium database (to 2009). Results of surveys in AW NRM in 2010 are not included. Nomenclature of all taxa in database, including old records, has been corrected and updated. NRM Region Alinytjara Wilurara Eyre Peninsula Kangaroo Island Northern & Yorke Adelaide & Mt Lofty Ranges SA Murray Darling Basin South East No. Species Recorded (to 2009) 200 741 752 656 747 116 673 No. Records in SA Herbarium database (to 2009) 409 8,771 9,104 5,834 7,854 219 5,981 It is also noted that the majority of the marine flora on reefs in the area are not endemic to the region. The statement occasionally quoted in summary documents that ―75% of the red algae in the GAB are endemic‖ is incorrect. It is noted that in a recent analysis of macroalgae records in the South Australian Herbarium, indicated that none of the South Australian endemic species of macroalgae in that data set were recorded in the AW NRM region (Baker and Gurgel, in prep.). Invertebrates Much of the standing biomass in the continental shelf waters of the GAB is likely to comprise the highly diverse attached fauna inhabiting the shelf, particularly sponges, bryozoans and ascidians. The species that make up these communities tend to vary from the inner to the outer shelf, and also decrease in abundance moving away from the coast (DEWR, 2006). Some of the invertebrate groups of significance in the GAB are discussed in more detail below. Sponges: Sponges are abundant and diverse in the Eucla Bioregion (of which the AW NRM marine region is part), but a complete species list does not exist. In 1995, during an Adelaide University survey (led by Dr Y. Bone) of the shelf waters deeper than 30m in the Great Australian Bight, 103 species of sponge were recorded, but many have not been described to species level (Appendix 2 in Sorokin et al., 2005). During a more recent SARDI survey of GAB Marine Park in 2001/02, around 100 species of sponge in 45 genera were recorded at 27 sampling points in the Eucla Bioregion (South Australian waters), and many of these have not yet been formally described (Sorokin et al., 2005, 2007). At one of the sites, near the Head of the Bight, 63 sponge species were recorded, including some currently undescribed species (Ward et al., 2003, Appendix 2). At least 63 (possibly up to 66) of the sponges recorded during 20 the first survey were not recorded during the second. An additional 5 or 6 species of sponge listed in the Australian Faunal Directory (Hooper, 2002 and 2005, in ABRS, 2009) as occurring in the GAB, were not recorded during either survey. From these 3 sources combined, and discounting joint occurrences, it is evident that more than 170 species of sponge occur along the GAB shelf alone (i.e. Eucla bioregion, waters less than 200m deep). The number of these sponge species that occurs in the AW NRM region has not been determined. Examples of sponges recorded during the SARDI survey in 2001/02 include species in the genera Acanthodendrilla, Acarnus, Anisocrella, Axinella, Callyspongia, Chalinula, Chondropsis, Clathria, Clathrina, Cliona, Coscinoderma, Cymbastela, Cribrochalina, Dactylia, Dictyodendrilla, Dysidea, Echinodictyum, Gelliodes, Halichondria, Haliclona, Hemitedania, Holopsamma, Iotrochota, Iotrochopsamma, Ircinia, Leiosella, Leucandra, Leucascus, Oceanapia, Phakellia, Phorbas, Psammocinia, Pyloderma, Raspailia, Reniochalina, Sigmaxinella, Siphonochalina, Stelleta, Suberites, Tabulocalyx, Tedania, Tethya, Trachycladus, Thorectandra, and Xenospongia. Sponge community composition changes across the shelf, with the highest diversity of sponges occurring in shallow inshore waters, declining with increased depth and distance offshore. The sponge fauna is dominated by species occurring in temperate Australia, with little affinity to fauna in tropical Australia (Sorokin et al., 2007). Unsurprisingly, comparisons of the deeper water sponge fauna of the GAB (i.e. 50m – 198m) with the shallow sponge fauna of the Investigator Group of islands in eastern GAB (Sorokin et al., 2007), show very little similarity between the two assemblages due to major differences in depth and habitat type. A number of chemicals isolated from sponges in the Great Australian Bight may have significant medical uses. For example, the widespread species Trachycladus laevispirulifer, contains organohalogens and other compounds which can kill leukaemia cells and also have anti-fungal activity (Vuong et al., 2001; Capon et al., 2008). Other isolates from sponges that occur in the GAB, such as the chemical hippospongins, are mildly antibiotic, and can inhibit the growth of Staphylococcus bacteria. After sponges die, the sponge spicules form part of the sediment in the Bight. 1.7.2 Bryozoans: This group is considered to be the main framework of the ecosystem on the cool-water sea-floor of the Great Australian Bight, providing environmental niches for other biota. Bryozoans also are an ―intermediate level‖ contributor in the food-chain; and are the major contributor to Tertiary limestone in South Australia (Bone, 1997). During SARDI surveys of the GAB Marine Park during 2002 and 2006, about 71 species of bryozoan were recorded in the shelf waters of the Eucla Bioregion. Although some of these were common and widespread species, such as Amathia tortuosa, Bugularia dissimilis and Cornucopina grandis, most were not identified to species level, and some may represent new species, still undescribed. Of the bryozoan species known from a wider area of the Bight, the number of these that occur in the AW NRM marine region has not been determined. The calcareous remains of bryozoans in the GAB, particularly in the outer shelf waters, form a significant component of the shelf sediment (Ward et al., 2003; Currie et al., 2008) and also the ―biogenic reef mounds‖ in the slope waters, which consist of diverse suites of bryozoans, together with coralline algae, echinoid spines, and benthic foraminifers, in a mudstone to packstone matrix (Feary et al., 2004). 21 Ascidians: The ascidian fauna of the AW NRM region has not been systematically investigated. However, to date, at least 130 ascidian species have been recorded in the South Australian section of the Great Australian Bight (Kott, 2006, in ABRS, 2009), all of these described and named by Kott. Of those, about 15 to 20 are known to date only from one locality where intense collecting occurred, mostly from islands in the Investigator Group, south-east of the Eucla Bioregion. One or two described species have been recorded to date only from the northern GAB (i.e. within the AW NRM region). Of the approximately 130 ascidian species known from a wider area of the Bight, the number of these that occur in the AW NRM marine region has not been determined. During the SARDI survey of the GAB Marine Park cited above, 75 different kinds of ascidian were recorded in the Eucla Bioregion (part of which includes the AW NRM region), but only a few of those have been identified to species level. It is noted that to date, many of the ascidians known from the Great Australian Bight are not endemic to the region, and the often misquoted ―high level of ascidian endemism within GAB‖ is not correct. Ascidians are a food source for various species of fish. Some molluscs, crabs, sea stars, nudibranchs, flatworms and other animals also feed on ascidians. Although recent surveys have provided some indication of the richness of the benthic fauna in the shelf waters of the GAB, much more research is required to document the distribution and biodiversity of the attached invertebrate fauna in this area, including the sponges, bryozoans and ascidians. Decapod Crustaceans: The decapod crustacean fauna of the AW NRM region has not been fully investigated, but a species inventory of crustaceans that occur in the South Australian section of the Great Australian Bight, indicates that at least 145 species of decapod crustaceans occur in the GAB shelf waters (Davie, 2002, in ABRS 2009; Poore, 2004; Currie et al., 2007), as shown in Appendix 1. Little is known of the decapod crustacean fauna in the Great Australian Bight, particularly the deeper shelf and slope waters. It is noted that during a beam trawl and benthic sled survey of the shelf and upper slope waters of the central and south-western Australian coast in 2005 (Poore et al., 2008), 76 species previously unrecorded south of Perth were found in the survey, and over the entire survey area, 175 species (33%) were new to science. The results are an example of how little is known about the decapod crustacean fauna of the continental margin of most of Australia (Poore et al., 2008), and the GAB is no exception. Some potentially limited range species of interest in the Great Australian Bight include the shallow subtidal eared sponge crab Haledromia bicavernosa (Zietz, 1888), which may be endemic within South Australia, and is known from the GAB and the South Australian gulfs (Davie, 2001, in ABRS 2009; O'Hara, 2002; Poore, 2004); (ii) the aesop prawn / aesop shrimp Periclimenes aesopius (Bate, 1863), a cleaner shrimp that is reportedly rare and known to date only from the GAB and gulfs region in South Australia (Poore, 2004; Gowlett-Holmes, 2008), and the swimmer crab Nectocarcinus sp., known to date from the outer shelf waters (140170m) in south-western WA and far western South Australia (Poore, 2004). This species resembles Nectocarcinus bullatus Balss, 1924 from Juan Fernandez Island off Chile, but is four times as big (P. Davie, pers. comm., cited by Poore, 2004). 22 Echinoderms: The echinoderm fauna of the AW NRM region has not been investigated, although a species inventory of echinoderms that occur in the South Australian section of the Great Australian Bight, indicates that at least 123 species of echinoderm occur in the area (Rowe and Gates, 1995, and O‘Hara, 2001, in ABRS, 2009), and these are listed in Appendix 2. This includes at least 37 species of sea star; 18 species of urchin; about 25 species of sea cucumber; at least 24 species of brittle star, and various basket stars and feather stars. Of the echinoderm species known from a wider area of the GAB, the number of these that occur specifically in the AW NRM marine region has not been determined. Species of note that may occur in the AW NRM marine region include Ophiothrix (Keystonea) hymenacantha H.L. Clark 1928 and Ophiothrix (Placophiothrix) albostriata H.L. Clark, 1928, both known to date only from type locality (Rowe and Gates, 1995, and O‘Hara, 2001, in ABRS, 2009). Another species of note is the endemic and live-bearing small sea star Parvulastra parvivipara (Keough and Dartnall, 1978), which is known from granite platforms in eastern Great Australian Bight (Roediger and Bolton, 2008). It is not known whether the distribution extends west as far as the Head of the Bight. Shelled Gastropod Molluscs: The gastropod fauna of the AW NRM region has not been investigated, although a species inventory of shelled gastropods that occur in the South Australian section of the Great Australian Bight, indicates that at least 485 species occur in the area (Baker, 2004 and references therein; Academy of Natural Sciences, 2006; OZCAM database, 2009; ABRS, 2009), and these are listed in Appendix 3. The central Great Australian Bight provides habitat for cowries, including those associated with sponges. Examples of cowries in the GAB area include apricot-coloured cowrie Umbilia (Umbilia) armeniaca Verco, 1912 (in deeper shelf waters, but possibly occurs at the outer edge of the AW NRM region); the cowrie Zoila venusta (including form profunda, recognised by collectors); humpback cowrie / black cowrie Zoila friendii thersites (including the pale form contraria, recognised by collectors), and Reeve‘s cowrie Austrocypraea reevei. Some of the species are rare / uncommon, and all temperate cowries are of conservation concern due to their vulnerable population characteristics (Ponder and Grayson, 1998; section 9.2 of Baker, 2004; Baker, in prep.). A number of other shells found in the region, such as the southern baler shell Melo miltonis, are considered particularly vulnerable due to limited distribution, large size, direct development of young, and over-collecting (Ponder and Grayson, 1998, cited by Baker, 2004). Other species of note that may be vulnerable due to their life history (e.g. direct development and limited dispersion), rarity, and/or to over-exploitation (from collecting) in the South Australian part of the range include sloping cowrie Notocypraea declivis; Cotton‘s volute Livonia nodiplicata; Günther‘s volute Nannamoria güntheri; desirable volute Amoria exoptanda; Flinders‘ vase shell Vasum (Altivasum) flindersi; punctate harp shell Austroharpa (Palamharpa) punctata, and the small white volute shell Notopeplum translucidum (Ponder and Grayson, 1998, cited by Baker, 2004). The large Jourdan‘s turban shell Turbo (Dinassovica) jourdani, which often occurs in tide pools, and amongst brown macroalgae in the shallow subtidal, has a limited distribution in both S.A. and W.A., and is considered vulnerable to population decline (Ponder and Grayson (1998, cited by Baker, 2004). Museum records show that at least two other shelled gastropod species of conservation concern may occur in the AW NRM: (i) Lesueur‘s sand triton Sassia (Cymatiella) eburnea lesueuri, a form of Sassia eburnea, for which there is a record from eastern GAB and this species might extend further west into the head of the GAB, and (ii) 23 Hexaplex conatus, a murex shell from S.A. and W.A., which is considered in the shell trading market to be rare. Some groups of shelled molluscs are species-rich in the shallow waters at the head of the GAB, including the top shells, marginella shells, turrid shells, triphorid shells and mitre shells (Appendix 3). Infauna: The infauna of the GAB shelf is species rich and abundant. In 2006, 65 samples of the macro-invertebrate infauna were collected by SARDI Aquatic Sciences from the shelf waters of the GAB (Currie et al., 2007), but most of these were in deeper waters outside of the AW NRM region boundary. Results indicated that (i) numbers of species and total abundance were typically highest near the Head of the Bight, where water temperatures are elevated, and also in inner-shelf waters off western Eyre Peninsula, which support high levels of plankton productivity; (ii) mobile, deposit-feeding organisms (primarily annelid worms and crustaceans) dominated the samples, and comprised over 25% of the abundance and 35% of the species collected over all samples (to the shelf edge); and (iii) the infaunal distribution pattern corresponds closely with spatial patterns in benthic species cover, and reinforces the notion that Head of the Bight is a ‗hotspot‘ of benthic biodiversity (as is western Eyre Peninsula) (Currie et al., 2007). More specifically, samples from the inner and mid shelf at the Head of the Bight during the 2006 survey were dominated by coarse grained sediments (average diameter 458.39 ± 29.51μm) with almost no mud (1-2%). The infauna in these sediments was species-rich (16 ± 2 species per 0.1 m2), and abundant (about 47 ± 15 individuals per 0.1 m2). During the 2006 survey, the fauna within the coarse sediments in the inner and mid shelf waters at the Head of the Bight comprised mainly crustaceans (48% of all major taxonomic groups recorded) and annelid worms (39%), with few molluscs (5%), chordates(2%) and echinoderms (2%). Minor groups include nemerteans (ribbon worms) (~ 1%), nematodes (round worms) (~ 1%) and sipunculids (peanut worms) (~ 0.5%) (Currie et al., 2007, 2009). During the SARDI survey, abundant infaunal species in the first 6 samples (closest to the seaward boundary of the AW NRM region) included the following (in descending order of abundance): the amphipod crustaceans Ceradocus rubromaculatus, Xenocheira fasciata, Gammaropsis persetosus, Caprella scaura, Dulichiella australis, and Leucothoe diemenensis; a tanaid crustacean in the genus Paratanais; the leptostracan crustacean Paranebalia longipes; the mussel Musculus nanus; the sipunculan worm Phascolosoma annulatum and the tanaid crustacean Kalliapseudes obtusifrons (Currie et al., 2007). These and other infaunal species collectively found in the inner and mid shelf waters (stations 1-6 during the 2006 SARDI survey) at the Head of the Bight are listed in Appendix 4. The ecological importance of GAB shelf infauna is indicated by the incidental and targeted catch of the Great Australian Bight (GAB) trawl fishery (see Part 2) which is principally comprised of latchet and deepwater flathead. Latchet feed directly on large infaunal species, while deepwater flathead are specialised for ambushing smaller species that forage for infauna along the bottom (DEWR, 2006). 24 Bony Fishes The Head of the Bight (and thus the AW NRM marine region) is in the middle of the Flindersian biogeographic province (Womersley and Edmunds, 1958), which runs from mid-western to south-eastern Australia. The fish fauna is a mix of temperate species typical of the region (many of those listed in Appendix 5), in addition to the following: (i) a number of subtropical fishes at the eastern edge of their range, which are more abundant in warmer waters of Western Australia. Examples include redlip morwong Cheilodactylus rubrolabiatus; yellow-spotted / brown-spotted boarfish Paristiopterus gallipavo; yellow-eyed red snapper Centroberyx australis; footballer sweep Neatypus obliquus; western wirrah Acanthistius serratus; smoothspine leatherjacket Cantheschenia longipinnis; blue-tailed leatherjacket Eubalichthys cyanoura; western black-spot pigfish Bodianus vulpinus; western form of Australian pineapplefish Cleidopus gloriamaris; smooth cocky gurnard Lepidotrigla sp.; two-spot eviota (goby) Eviota bimaculata; western red-banded grubfish Parapercis naevosa; rigid boxfish Caprichthys gymnura and western smooth boxfish Anoplocapros amygdaloides; and (ii) a number of cooler water, south-eastern Australian fishes at the western edge of the range, and examples include blue warehou Seriolella brama; Derwent flounder Taratretis derwentensis; dragonet Bovichtus angustifrons; southern cardinalfish Vincentia conspersa; little conger Gnathophis longicauda; pot-bellied seahorse Hippocampus bleekeri / abdominalis; short-headed seahorse Hippocampus breviceps; barred toadfish Contusus richei, and common shore-eel Alabes dorsalis. A few fish species have their centre of distribution in the shallow shelf waters of the GAB, and are not known outside of South Australia and south-western Australia. Examples include the eelblenny Peronedys anguillaris and shortfin snakeblenny Ophiclinus brevipinnis. The crested threefin Trinorfolkia cristata, an apparently endemic species, confirmed to date from numerous locations in South Australia and no other State (Baker 2009), occurs in the shallow waters of the GAB. In the GAB, the strength and timing of the Leeuwin Current can affect abundance of some fish species in South Australia, such as southern bluefin tuna, blue mackerel, horse mackerel, west Australian salmon, and Australian herring (tommy ruff) (Lenanton et al., 1991; Pearce and Caputi, 1994; uncited reference, in Director of National Parks, 2005). The distribution and abundance of these species is influenced by the seasonality, strength and timing of the Current. Tropical and sub-tropical pelagic fishes sometimes occur in the shelf waters of the GAB, due to the influence of the easterly moving Leeuwin current, which varies inter-annually in strength. Appendix 5 lists the bony fish species that have been recorded in the AW NRM region, or are highly likely to occur there, based on geographic distribution, habitat and depth range. For many of the 332 species listed in Appendix 5, the Great Australian Bight has been included in a generalised distribution map of the species range (e.g. Gomon et al., 2008), or there are records from the area (e.g. S.A. Museum, W.A. Museum, Museum Victoria records, in OZCAM database, 2009). However, a number of shallow water species that have been recorded in both south-eastern and south-western Australia have been excluded from the table, due to lack of records from the Great Australian Bight, but it is noted that such species might also occur in the AW NRM region. Examples include the weedfishes Cristiceps aurantiacus Castelnau 1879, which occurs in kelp beds, and Heteroclinus sp. 2, found in low macroalgae on coastal reefs; 25 the flathead Platycephalus longispinis Macleay 1884; the reef-dwelling red indianfish Pataecus fronto Richardson 1844; the tiger pipefish Filicampus tigris (Castelnau, 1879); the mullet Mugil cephalus Linnaeus 1758 and the silver Fish Leptatherina presbyteroides (Richardson, 1843). South-eastern species, for which the eastern GAB area (e.g. Streaky Bay, or Ceduna) is the western limit have also been excluded. Examples include the blenny Parablennius tasmanianus (Richardson 1842) and the spotted flounder Ammotretis lituratus (Richardson, 1844). Nevertheless, Appendix 5 indicates that at least 332 species of bony fishes are highly likely to occur in the AW NRM marine region, these being a mix of resident reef fishes, sand-dwelling and seagrass-dwelling fishes, pelagic schooling fishes that live in the water column, migratory pelagic fishes pass through the area, and oceanic fishes that are occasionally stranded inshore. Many factors likely contribute to this species richness, including the mix of inshore and shelf habitats (see above), and the biogeographic position of the region. A number of potentially threatened bony fishes occur in the AW NRM region. Some examples include southern bluefin tuna, mulloway, western blue groper, harlequin fish, and Southern Blue Morwong, amongst others. The conservation status of these species, and potential threats in the GAB, are discussed in section 3.1.1 Fishing in Part 3. Cartilaginous Fishes The Great Australian Bight is a significant area for the protected white shark Carcharodon carcharias. The White Shark is a relatively long-lived (more than 30 years), late-maturing species, with low reproductive output every two to three years (Klimley and Ainley, 1996; Bruce et al. 2001, cited by Baker et al., 2008). Females grow larger than males, and are not reproductively active until they are over 5 m long and about 18 years old (Bruce et al., 2001). White Sharks are top order predators, and are thus rarely preyed upon, other than by even larger carnivores such as killer whales Orcinus orca (Pyle et al., 1999). Juvenile White Sharks (under 2m) are commonly encountered in the inshore waters of the Great Australian Bight. The relative abundance of small White Sharks in the GAB may be due to pupping in this area or nearby. Movements of White Sharks are known to increase seasonally at the Head of the Bight, and may be linked to the seasonal availability and movements of prey, including snapper, mulloway, small gummy shark s and Australian salmon, and to the calving of southern right Whales. White sharks may also seasonally visit Australian sea lion colonies in the GAB. White sharks are highly migratory. Tagging of white sharks by CSIRO at the Neptune Islands off southern Eyre Peninsula has shown that some White Sharks move across the GAB into Western Australia. One female that was tagged in 2006 moved 3,800 kilometres from Port Lincoln in S.A. to Exmouth Gulf in north-western W.A., passing through the deeper shelf waters of the Bight along the way (CSIRO, 2008a). From W.A., the shark then began travelling back towards the GAB. Another white shark, a male tagged in 2004, also travelled to south-western WA, and one of the only points close to the coast where it passed during the journey was the S.A. / W.A. border (CSIRO, 2008b). Another male white shark tagged in 2004, moved northwestwards along the western Eyre Peninsula coast, into relative shallow waters at the very Head of the Bight, then travelled back down to islands at the bottom of Eyre Peninsula (CSIRO, 2005). The GAB is an important area for the school shark, a long lived species of ground shark (more than 50 years) which occurs all over South Australia and southern Australia, and in numerous 26 other countries. school sharks are highly migratory. In Australia, tagging has shown mixing across most of the southern half of the continent (with movements of more than 2,000km across southern Australia) and more than 30 tagged animals have moved across the Tasman Sea between Australia and New Zealand (and vice versa), with movements of more than 3,500km (Olsen, 1984; Coutin et al., 1992; Stevens and West, 1997, cited by Bruce et al., 2002; Hurst et al., 1999; Walker et al., 1999, 2000; AFFA, 2002; Brown et al., 2000, cited by Cavanagh et al., 2003). In addition to long distance migration, more recent archival tagging suggests that complex and geographically distinct movements also occur in southern Australia. During summer, archival tagged school sharks favoured the outer GAB and Tasmanian east coast, whereas the sharks in those regions moved to Kangaroo Island and Tasmanian west coast during winter. Movements occur over a wide time period in South Australia (West and Stevens, 2001, cited by Bruce et al., 2002). Migration is linked to feeding and reproduction (Last and Stevens 1994, cited by Pogonoski et al. 2002; Bruce et al., 2002). In late summer and winter, schools of adult sharks move either to deeper waters at the edge of the continental shelf in the Bass Strait region, or to warm waters of South Australia and New South Wales. At the edge of the continental shelf, reproduction occurs. Adult school sharks then travel southwards and shorewards in the spring to converge along the coastlines, where they feed in schools that vary in composition of individuals. About half of all adult females in these schools may be pregnant during the breeding season, and these visit the pupping grounds to renew the cycle (Compagno, 1984, cited by FAO, 2000). There is a hypothesis that school sharks may travel ―home‖ to their natal mating and pupping grounds (Walker and Punt, 1998, cited by Bruce et al., 2002). In the GAB, fishing records indicate the presence of school sharks in deeper waters south-west of Ceduna and Streaky Bay, and around the Nuyts Archipelago; also Head of the GAB; deeper waters throughout the entire GAB (both continental shelf and slope waters); and waters near W.A. / S.A. border. At the Head of the Bight, school sharks are strongly associated with many reef patches, also an ancient submerged coastline. s travel into the shallow waters at the northern end of the Bight to give birth to pups. Generally, school sharks appear to have fairly discrete pupping and nursery areas, often in shallow, protected bays and estuaries, but also in coastal / ocean beach habitats (Olsen, 1954; Stevens and West, 1997; Stevens, 1999, cited by Bruce et al., 2002; Walker et al., in Cavanagh et al., 2003). One area considered to be significant for pregnant female school sharks is the Head of Great Australian Bight – Eyre Bluff to the Western Australian border. It is thought that pregnant females from south-eastern Australia may travel to South Australia, aggregate in particular areas, then migrate back to pupping areas in Tasmania (Walker et al., 1989 and 2000; Prince, 1996, cited by Bruce et al., 2002; AFMA, 2003). The Head of Bight aggregation area includes some of the GAB Marine Park (AFMA, 2003). There are 5 other areas identified as being significant for pregnant female school sharks and/or pups in South Australia, and Head of Great Australian Bight to Fowlers Bay is one of those areas (AFMA, 2002), which includes a significant portion of the AW NRM region. Although the movements of school sharks have not been studied in the GAB, it is noted that in south-eastern Australia, pregnant females move into shallow, partly enclosed bays and estuaries in late spring and early summer, and depart to offshore feeding grounds after dropping off their young. Most young-of-the-year depart the pupping grounds in late summer and move offshore, but mostly return to the bays (and estuaries, in some areas) of their birth the following spring; some juveniles may switch to adjacent bays. Some juvenile school sharks may remain in shallow protected waters for up to two years before moving offshore. Juveniles of 27 about two years old join schools of immature sharks that are inshore or offshore along the coast (data by Walker et al., cited by FAO, 2000; Bester 2003). Another ground shark, the gummy shark Mustelus antarcticus, has a similar distribution to the school shark in southern Australia, and is abundant in the GAB. gummy shark s are often associated with soft seafloor sediments (Bax and Williams, 2001), and their diet is dominated by squid and octopus, crustaceans (with crabs mainly taken by small gummy sharks, and lobsters by larger ones), and bony fishes, including toadfishes (Simpfendorfer et al., 2001). gummy shark s are moderately long-lived (to about 16) and fecund (up to 38 pups) (Lenanton et al., 1990; Moulton et al., 1992; Walker, 1996, 2007). Like school sharks, gummy shark s visit the shallow protected waters of the northern GAB to give birth to pups. Whiskery shark Furgaleus macki and sawsharks (common sawshark Pristiophorus cirratus and /or southern sawshark Pristiophorus nudipinnis) also occur in the GAB, including shallow waters. In the GAB, the whiskery shark is recorded over reef patches in the mid to outer shelf (100-200m), but is also found inshore. Generally, it occurs on or near the sea floor, commonly near rocky bottom, beds of kelp or other macroalgae, or seagrass beds (Compagno, 1984; Simpfendorfer and McAuley, 2003). The diet of whiskery sharks is highly specialised, with cephalopods (mainly octopus and squid) making up approximately 95% of food eaten (Simpfendorfer et al., 2001). Bony fishes, and crustaceans such as lobsters, are a minor part of the diet (Kailola et al., 1993; Last and Stevens, 1994). There is a seasonal reproductive cycle. Mating is most likely to occur from August to September, with females storing spermatozoa until ovulation in late January to early April. Gestation lasts 7–9 months, with parturition from August to October. Litter sizes vary from 4 to 28 (mean 19). In a study off south-western W.A. (based on gill-net fishery specimens), there was a significant linear relationship between litter size and maternal length. Mature males mate each year, but females produce litters every second year (Simpfendorfer and Unsworth, 1998). There is little information on movements with age and size, but it is likely that juvenile whiskery sharks occur inshore. Similarly, there is little information on the distribution or relative abundance of sawsharks in the GAB, but abundance is assumed to be low, given the low bycatch in gillnet, trawls and long-lines. In South Australia, reported sightings and captures of whaler sharks (mainly bronze whaler C. brachyurus, but also black whaler / dusky shark C. obscurus in some areas) occur throughout coastal waters, ranging from high energy surf beaches and off rocky coastal headlands, to more sheltered waters of the upper gulfs (Cappo, 1992, cited by Jones, 2008). The bronze whaler Carcharhinus brachyurus is found throughout the AW NRM region and surrounds, including inshore areas near Eucla and Head of the Bight, and also outside bays to the east of the AW NRM region. According to commercial fishing data, most records in the area are from shallow waters, extending to about 100m deep. Globally, there is evidence for both migration and strong site association in this species (Cliff and Dudley, 1992), and data from South Australia also support this. Based on a small number of tag recaptures, there may be some seasonal movement in South Australia (e.g. away from colder upper gulf waters in winter, into warmer offshore waters in the GAB and off southern Kangaroo Island), but more data are required (Jones, 2008). One bronze whaler that was tagged in the gulf waters during summer, moved to waters north-west of Pearson Island in the eastern GAB by winter (Jones, 2008). Conversely, tagged adult bronze whalers in South Australia have also been re-sighted at their tagging location after a year at liberty, suggesting this species, like others in the family Carcharhinidae, is philopatric (i.e. faithful to place of birth, through residency or periodic return) (I. Gordon, pers. 28 obs., cited by Duffy and Gordon, 2003). Over the broad geographic range, C. brachyurus occur singly and in loose schools, sometimes numbering hundreds of individuals (Smale, 1991; Cappo, 1992; Cliff and Dudley, 1992; C. Duffy, unpubl. data, cited by Duffy and Gordon, 2003). In some parts of the range, adult males are present in subtropical regions throughout the year, whereas females and immature C. brachyurus migrate into these regions during winter. Most adult females return to temperate regions to breed in the spring. Males also appear to migrate to higher latitudes in late winter-spring, presumably to mate (Cliff and Dudley, 1992). Mating probably occurs offshore as females disperse from the nursery areas. Juvenile and adult C. brachyurus segregate by size and sex (Muñoz-Chápuli 1984; Smale, 1991; Cliff and Dudley, 1992; Chiaramonte, 1998, cited by Duffy and Gordon, 2003). Juveniles occur in shallow water (less than 30m depth) year-round, whereas adults are most abundant inshore during spring and summer. Adults and sub-adults are found over the shelf and around offshore islands and banks throughout the year. The movement of adult females inshore in spring is related to breeding. In Western Australia, females move inshore to drop their young (Harrison, 2001). Nursery areas tend to be large and ill-defined, but can include shallow banks and large shallow bays, as well as the open coast (Muñoz-Chápuli, 1984; Smale, 1991; Cappo, 1992; Chiaramonte, 1998; Fergusson and Compagno, 1995, cited by Duffy and Gordon, 2003). Segregation of bronze whalers also occurs along latitudinal gradients (Muñoz-Chápuli, 1984; Smale, 1991; Cliff and Dudley, 1992). C. brachyurus inhabiting temperate areas can range into higher latitudes during the summer (Ayling and Cox, 1982; Compagno, 1984; Compagno et al., 1989: Cappo, 1992: Last and Stevens, 1994; Chiaramonte, 1998). These latitudinal movements may be in response to changes in water temperature or prey migrations (Compagno, 1984; Compagno et al., 1989; Cliff and Dudley 1992; Chiaramonte, 1998). This species eats many types of bottom-living and pelagic cephalopods and fishes (see Duffy and Gordon, 2003). Food items in South Australia include schooling Australian salmon (a main dietary item), rock flathead and snook (Cappo, 1992; Jones, 2008). The fishery for bronze whaler sharks is discussed in Part 2 of this report, and threatening processes are discussed in Part 3. Also occurring in the GAB is the dusky shark Carcharhinus obscurus, which is found over a broader depth range (from inshore waters to the upper continental slope) (Last and Stevens, 1994; Daley et al., 2002). The abundance of C. obscurus in S.A. is much lower than in W.A. (McAuley, 2005, cited by Jones, 2008). In temperate and subtropical areas the species is migratory, moving north during the warmer summer months and retreating south when water temperatures drop (Musick et al., 2007). In Western Australia, a tagging study showed that there was some movement into South Australia of tagged individuals (including east of S.A. / W.A. border, and also a small number into the S.A. gulfs), but main seasonal movements were north and south along the west coast of W.A. (Simpfendorfer et al., 1999; McAuley, 2005; B. Hay, pers. comm., cited by Jones, 2008). There are potentially distinct nursery areas in inshore waters. In Australian waters, adolescents and adults appear to move inshore into shallower water (less than 80m depth) off WA during summer and autumn, with neonates occupying separate inshore areas (Last and Stevens, 1994). In Australia, Dusky Sharks may breed in North West Shelf waters during winter, and migrate southwards, giving birth to their young off the south-west off W.A. during autumn and winter (McAuley, pers. comm., cited by Pogonoski et al., 2002; McAuley, 2005). Despite a small number of tagged juvenile Dusky Sharks being recaptured in South Australia, most young probably remain in the southern waters of W.A. for a number of years before migrating northwards, to become part of the mature stock, which is concentrated in W.A. (McAuley, pers. comm., cited by Pogonoski et al., 2002; McAuley, 2005). 29 The diet of C. obscurus includes a variety of reef fishes, bottom-dwelling fishes, and pelagic bony fishes, along with other shark species, rays, crustaceans, cephalopods, barnacles, and whale meat. In W.A., the main diet is pelagic fish such as Sardinops sagax, octopus and squid, and cartilaginous fishes become a more important part of the diet in larger, older sharks (N.B. cannibalism also occurs) (Simpfendorfer et al., 2001). The GAB provides an important habitat for the southern banded wobbegong / large ornate wobbegong Orectolobus halei, which is caught in the trawl bycatch in shelf waters. In the GAB, this species may be concentrated along the mid shelf and shelf edge at certain locations, but also in inshore reefs at the head of the GAB. Research in south-eastern Australia has shown that temperate wobbegongs favour areas of ―high topographic complexity and crevice volume‖ (Carraro and Gladstone, 2006). Wobbegongs are usually nocturnal sharks, that rest on the bottom during the day in caves, under ledges on reefs, and in trenches, and seek prey in reef habitat at night. Prey includes bottom-dwelling fishes, small sharks and rays; octopuses and other cephalopods; crayfish / lobsters, crabs and other crustaceans (probably taken by juveniles) (Compagno, 2001, 2005; Last and Stevens, 1994; Australian Museum, 2003; Compagno, 2005; C. Huveneers, pers. comm., 2007). Fishes recorded in the diet of O. halei include yellowtail scad Trachurus novaezelandiae, mulloway Argyrosomus japonicas, western blue groper Achoerodus gouldii, drummers (Kyphosus species) and various scombrids (mackerels, tunas etc) (Huveneers, 2007). Abalone has also been recorded in the diet (Kailola et al., 1993). Wobbegongs occur singly, or often in aggregations during the day (Pollard et al., in Cavanagh et al., 2003), sometimes with several animals piled on top of one another (Compagno, 2005), making them particularly vulnerable to trawl capture. Another benthic shark species in the area is the Port Jackson shark Heterodontus portusjacksoni, widespread throughout southern Australia, and commonly found on rocky reefs, on sandy and muddy bottoms, and in seagrass beds. Port Jackson sharks are fished from the coast in summer, especially around Coombra lagoon (A. Loisier, AW NRM Board, pers. comm., 2010). The species is relatively slow growing and long-lived (to 40 years), with delayed maturity (6-8 years in males, and 8-11 years in females) (McLaughlin and O‘Gower, 1971; Rodda, 2000; Izzo, 2006, cited by Baker et al., 2008). Adult Port Jackson sharks feed mainly on bony fishes, decapod crustaceans, squid, octopus and echinoderms, whereas juveniles mostly eat worms (Powter, 2007; Svane et al., 2007). Wobbegongs prey on Port Jackson sharks, as well as many other items (see previous paragraph) (Compagno, 2001; Huveneers, 2006). Although relatively long-range (hundreds of km) migrations have been recorded in south-eastern Australia, studies of this species in South Australian gulfs indicate much smaller scale movements, mostly less than 20km (Svane et al., 2007). In both south-eastern Australia and South Australia, Port Jackson sharks have been recorded in spawning aggregations, and animals may return annually to the same areas to lay eggs. Port Jackson sharks also use a suite of specific ―resting sites‖ during non-breeding periods. When the sharks at one site are disturbed, they will move directly to another of their resting sites. After the breeding season, males move into deeper waters followed by the females in late September or October. The remarkable homing ability of these sharks, which can return to the same site after several years of migration, has given rise to a hypothesised spatial memory in these sharks (O‘Gower 1995). The inner and mid shelf areas of the Head of the GAB (including waters of the AW NRM region) are also important for the smooth hammerhead Sphyrna zygaena, which is concentrated in areas of heavy reef and numerous reef patches, and also sandy areas, particularly shallow to 30 mid shelf waters at the Head of the Bight (and also western Eyre Peninsula). Over the broad range of this species on the shelf and upper slope, the preferred depth is close to shore. Charter boat captures from the eastern Great Australian Bight indicate seasonal aggregation in shallow waters. In other parts of the range (outside of Australia), juveniles may be abundant and form large aggregations, and various nursery areas have been identified. Adults occur either singly or in small groups. This species often occurs at the surface in the open ocean, and can form enormous schools during migration to cooler latitudes during the summer months. However, the relative abundance, seasonal movements and age- and reproduction-related movements of smooth hammerhead in the GAB have not been investigated. Smooth hammerhead is primarily a piscivore, and feeds on a variety of bony fishes including clupeids (sardine and herring family) and small scombrids (tuna and mackerel family). This shark species also eats smaller sharks (as well as its own species), stingrays, squid and other cephalopods, and benthic crustaceans (Stevens, 1984; Smale, 1991; Bornatowski et al., 2007). Some of the shark species in the vicinity of the Head of the GAB are known mainly from offshore areas at the edge of the shelf, including the migratory shortfin mako Isurus oxyrinchus, and the sharpnose sevengill shark Heptranchias perlo and broadnose sevengill shark Notorynchus cepedianus. Appendix 6 lists the cartilaginous fish species that have been recorded in the AW NRM region, or are highly likely to occur there, based on geographic distribution, habitat and depth range. South-eastern species, for which the eastern GAB area (e.g. Investigator Group, Streaky Bay, or Ceduna) is the western limit have been excluded. Examples include the banded stingaree Urolophus cruciatus. Species known mainly from outer shelf and slope waters of the GAB are also excluded (e.g. wide stingaree Urolophus expansus and Bight skate Dipturus gudgeri. Some of the species occurring in the GAB are considered vulnerable at international level and have been listed in the IUCN‘s Red List (see Appendix 6). Marine Mammals Southern Right Whale Eubalaena australis: The Head of the Great Australian Bight is a significant area for southern right whales. Although numbers of southern right whales are now increasing at 7-8% per year (DEH South Australia, 2006), this whale is still considered a threatened species under national and State level legislation, despite a recent downgrading of threat category to least concern on a global scale, by the IUCN. Southern right whale is listed under the Commonwealth Environment Protection and Biodiversity Conservation Act 1999 (EPBC Act) as endangered. It is classified under the South Australian National Parks and Wildlife Act 1972 as vulnerable. Around 1,500 of the 12,000 or so southern right whales remaining worldwide visit Australia during the austral winter, usually arriving in May after their journey northwards from Antarctica, and departing around October. The global number is still a small percentage of the estimated 60,000 - 100,000 that existed prior to whaling during the 19th century (DEH South Australia, 2006). Approximately 26,000 southern right whales were reported to have been killed through whaling in Australian and New Zealand waters from 1820 to 1850 alone. Fowlers Bay was the closest whaling station to the GAB, during the 19th century, and whales still pass through the area today. The Head of Bight is a critical gathering area for southern right whales, with up to half of the Australian population (10% of the global population) using the area for mating (southern right whales are mature at about 7 years old), calving and 31 socialising. The Head of the GAB is one of the three most important calving grounds on the Australian coast, with 25-45 calves born each year, and more than 100 whales per day may be present in peak periods of peak years (Pirzl and Burnell, 2004; Pirzl, 2008; A. Loisier, AW NRM Board, pers. comm., 2010). During the past 2 years to 2010, more than 100 (and up to 140) whales were counted on some days towards the end of the season (A. Loisier, AW NRM Board, pers. comm., 2010). Mark-recapture estimates show that about half the Australian population of this species uses the site (Pirzl and Burnell, 2004). Whales gradually enter the nursery waters of the Bight in early June, where those that are pregnant give birth. During autumn and early winter, mostly adults are observed, but by late winter and spring, mothers with calves are commonly seen swimming along the cliffs of the GAB. At such time, there are often 70 or more whales visible from a viewing platform in the area. Each year between 25 and 55 calves are born at or near the Head of Bight, making it one of the three largest breeding areas along the Australian coast (DEH South Australia, 2006), and one of the most significant nursery areas (Pirzl, 2008). Each day, mothers and new born calves nurse, rest, play and travel. Mothers and calves tend to spend more time at the Head of Bight than lone whales, which tend to leave and return over the season. The average residence of females that calve at the Head of the Bight is 71 days, higher than the average residence time of unaccompanied adults (20 days) (Burnell and Pirzl, 2006; Pirzl, 2008). A study at the Head of the Bight recorded 20 calving females in 2002; 28 in 2003 and 32 in 2004 (Pirzl, 2008). Southern right whales calve about every 3.64 ± 0.13 years at Head of Bight (Burnell, 2001, cited by Pirzl, 2008). Whales in the Head of Bight area often occur about 500m from shore (Pirzl, 2008). During the austral summer, whales head back south, travelling and feeding well offshore in the southern Ocean. During the early 2000s, a genetics project which utilised skin biopsy samples showed that genetic differences (in term of mitochondrial DNA) exist between southern right whales in the GAB, and those in Western Australia (research by N. Patenaude and R. Harcourt, Macquarie University, 2001-04). The genetic research has indicated maternally-directed substructure within the Australian population (Patenaude and Harcourt, 2006) and long-term distribution changes may be the result of extirpation of localised sub-groups (Pirzl, 2008). During the early to mid 2000s, a project on habitat selection and usage by southern right whales at the Head of Bight and two other calving areas, showed that the whales exhibit significant spatial structure at a fine-scale (within calving grounds), and that the location of individuals is primarily determined by the location of other whales (i.e. spatially auto-correlated) (Pirzl, 2008). Depth apparently has a big influence on distribution of the whales, with a general preference for shallower water (particularly by mothers with young calves) being modulated by specific population factors and local environment conditions. At the Head of Bight, where whales aggregate seasonally, about 29% of the area from shore to 3km seawards is less than or equal to 10m deep, and 60% is 20m deep or less (Pirzl, 2008). A long-shore distribution pattern was also evident for the whales in this area, and during the season, whales tend to move to deeper, more westerly waters of the Bight as calves grow larger / older (Pirzl, 2008). Australian Sea Lion Neophoca cinerea: Populations of Australian sea lions in the GAB, which were likely never commercially harvested in the past, provide a very important genetic and geographic bridging population between the South Australian and Western Australian sea lion populations (Dennis and Shaughnessy, 1996). Previously, in 1994, a survey of potential haulout sites of the Australian sea lion along the coastline of the Great Australian Bight (from Twin Rocks to Wilson Bluff: distance = 206km) reported 289 sea lions at 23 sites, widely dispersed at the base of the Bunda Cliffs. The animals were hauled out on perched platforms formed by collapsed sections of cliff at various levels above the sea. Of these, 37 sea lions were recorded 32 in a deep cave accessed from the sea. The total included 86 pups aged under 12 months, which were considered to probably have been born in the region (Dennis and Shaughnessy, 1996). The Australian sea lion sites located in 1994 were surveyed again in August-September 1995, during a predicted breeding season. In that survey, a total of 284 sea lions was recorded at nine sites in South Australia and one site in Western Australia (2km west of Twilight Cove). The 1995 survey included 90 pups under six months of age. Overall, 10 breeding sites and 14 haul out sites were recorded. By extrapolating from the number of sea lion pups found in 1994, the population for the Great Australian Bight region in South Australia was estimated to be between 613 and 774 during the 1990s (Dennis and Shaughnessy, 1996), which represented approximately 9.3% of the South Australian population or 6.6% of the total world population for this species at the time (Dennis and Shaughnessy 1996). At least 7 breeding and haul-out sites are known along the Head of Bight area (SARDI Aquatic Sciences, 2006), known as Bunda Cliffs B1, B2, B3, B5, B6, B8 and B9 (Goldsworthy and Page, 2009; Goldsworthy et al., 2009b). The seven GAB populations are spread from the Head of the Bight westwards towards the W.A. border, and there is an eighth colony at western Nuyts Reef (Goldsworthy, 2009b). Recent estimates of pup numbers at the Bunda Cliffs sites are as follows: B1 = 15 , B2 = 5, B3 = 31, B5 = 43, B6 = 12, B8 = 38 and B9 = 17 (Goldsworthy et al., 2009a). This species has also been recorded in waters adjacent to Wahgunyah Conservation Park, east of the AW NRM region (Caton et al., 2007). Foraging area maps are being developed for Australian sea lions, and it is noted that adult male sea lions utilise the entire continental shelf to forage, where they overlap with adult females, but the adult males also forage further out to sea, in deeper waters than the females (Goldsworthy and Page, 2009). The relatively small populations of females scattered through the GAB may be faithful to their natal colonies, as was shown in populations in Western Australia (Campbell, 2003, cited by DEWR, 2006). An extended rearing period for older pups suggests that learning about the location of local feeding grounds is important to the species. The breeding cycle is about 13-14 months, and colonies in different parts of the GAB cycle have staggered breeding times, perhaps to help spread their predation pressure on limited shared resources more evenly through the seasonal cycle. It could also provide further protection for small genetic populations against predation by making it more difficult for predators to take advantage of young pups during predictable pupping seasons. Australian sea lions in the GAB are noted for being relatively thin, in comparison to those foraging in the more productive waters off Kangaroo Island to the east (uncited reference, in DEWR, 2006). New Zealand Fur Seal: During the 1994 survey of Dennis and Shaughnessy, 12 New Zealand fur seals Arctocephalus forsteri were sighted along the coastline of the Great Australian Bight from Twin Rocks to Wilson Bluff (Dennis and Shaughnessy, 1996). Goldsworthy and Page (2009) recorded only one breeding site, at west Nuyts Reef. The colony in the Nuyts Reef area may be the most westerly of the breeding colonies in South Australia (Shaughnessy et al., 1994, 2005; Director of National Parks, 2005). Other Species: Appendix 7 lists the marine mammal species recorded in the Great Australian Bight, and their conservation status at international, national and South Australian levels. Many of these mammal species occur only rarely, as strandings of dead or injured animals (e.g. Ling, 1991; Kemper and Ling, 1991; Director of National Parks, Commonwealth DEH, 2005). Beaked whales, including such species as Gray‘s beaked whale Mesoplodon grayi and straptooth whale Mesoplodon layardii, are occasionally stranded at the Head of the Bight (e.g. DEWHA, 2010). A number of species are regularly sighted in situ. For example, the killer whale Orcinus orca, has been recorded just east of the AW NRM region, approx. 2km south of Fowlers Bay (1989), and 33 at several sites further east, in the eastern GAB (Ling, 1991). There may also be records from the Head of the Bight Area, as evidenced from the Southern Oceans Orca Database, held at Deakin University (Morrice, undated). The presence of killer whales may be related to the abundance of pinnipeds in the region, upon which they feed (Kemper and Ling 1991, cited by Director of National Parks, 2005). Humpback whales are commonly seen from the cliff tops at the Head of the Bight, and this fast-moving species usually travels further offshore compared with southern right whales (A. Loisier, AW NRM Board, pers. comm., 2010). Some species, such as blue whales and sperm whales travel seasonally through the Bight, but these species are not commonly recorded inshore. Dolphin species which are common wide-ranging throughout South Australia, including the GAB, are Indo-Pacific bottlenose dolphin Tursiops aduncus, often in estuarine and coastal waters; common bottlenose dolphin Tursiops truncatus, in coastal and offshore waters, and the common short-beaked dolphin Delphinus delphis (Kemper, 1998, 2004). Marine Reptiles Reptiles such as the green turtle Chelonia midas are occasionally recorded in the GAB, with museum records from the Ceduna area, and in situ sightings from further west, including the Nuyts Archipelago and waters further west, in the northern GAB. The green turtle is classified internationally as endangered with a decreasing population trend (IUCN, 2009), and is listed under the Commonwealth EPBC Act 1999 as vulnerable in Australia. There are also records of the loggerhead turtle Caretta caretta in the GAB area, and at locations further east, such as the SA gulfs region. The leathery (or leatherback) turtle Dermochelys coriacea is also recorded occasionally in the GAB (e.g. Nuyts Archipelago), and the GAB may be an important feeding area (DSE, 2009). The leathery turtle is classified internationally as critically endangered and the loggerhead as endangered (IUCN, 2009), and both the leathery and the loggerhead are also listed under the Commonwealth EPBC Act 1999 as endangered in Australia. The occasional presence of migratory turtles in the GAB may be linked to transport eastwards in the Leeuwin Current from western Australia. Turtles occasionally wash up dead in the eastern GAB, in areas such as Anxious Bay (e.g. Edyvane et al., 2004). All marine turtle species occurring in South Australian waters are also listed under the South Australian National Parks and Wildlife Act 1972. Coastal Birds and Sea Birds According to maps in Pizzey (1988), about 60 species of coastal and marine birds are likely to occur in vicinity of the head of the Great Australian Bight. During a recent survey of the Alinytjara Wilurara lands in September 2008 (Dennis, 2008), 28 bird species of coastal and/or marine association were recorded, several of which are listed under South Australian and/or Commonwealth legislation as threatened species, and others of which are listed migratory species under international conventions, such as JAMBA and CAMBA. These, and other marine bird species recorded either regularly or irregularly in the GAB from previous surveys (e.g. Carpenter, 2005, cited by Caton et al., 2007), are listed in Appendix 8, which also lists the conservation status of these species at international level (IUCN; JAMBA, CAMBA and 34 ROKAMBA); national level (Commonwealth EPBC Act 1999) and State level (South Australian National Parks and Wildlife Act 1972). Some species are seasonal, such as the migratory sanderling and red-necked stint (both present during summer), and the southern giant-petrel (present during winter). The endangered white-bellied sea eagle Haliaeetus leucogaster has been recorded in the AW NRM region, including the Head of the Bight, and sections of the Nullarbor cliffs west of the Head of the Bight (Caton et al., 2007). White-bellied sea eagles nest along cliffs, rock pinnacles, escarpments and tall trees. The number of territories currently occupied by the white-bellied sea eagle is reported to have declined in the far west of South Australia (Dennis and Lashmar, 1996, cited by Caton et al., 2007). Dennis (2008) recorded an active nest of H. leucogaster in a previously unrecorded territory, in the eastern section of the Bunda Cliffs. The white-bellied sea eagle is very sparsely distributed in the region. Surveys of Dennis (2008) have indicated that there may be as few as 3 individuals in the broad region from Laura Bay (eastern GAB), through to the S.A. / W.A. border, hence the number within the AW NRM region is very low. The endangered osprey Pandion haliaetus nests on rocky headlands, stacks, cliffs, live or dead trees and artificial platforms. Several active osprey sites have been identified in the northern Great Australian Bight and Yalata (T. Dennis pers. comm., cited by Caton et al., 2007). Examples including the Head of the Bight, and a site near the Merdayerrah Sandpatch (Dennis, 2008). Like the white-bellied sea eagle, the osprey is sparsely distributed in the region, and the surveys of Dennis (2008) have indicated that there may be as few as 8 osprey in the broad region from Laura Bay (eastern GAB), through to the S.A. / W.A. border (Dennis 2008). Therefore, the number in the AW NRM region is even lower. Parts of the AW NRM region provide significant habitat for the vulnerable hooded plover Thinornis rubricollis rubricollis. Examples of significant habitats include the beach and foredunes in the vicinity of the Dog Fence beach (northern end of Wahgunyah Conservation Park, near the eastern border of the AW NRM region), with one active nest recorded recently 1km south-east of Tjitji Tjutaku campsite (Dennis, 2008). Other locations where this species might be present include the coast adjacent to Wahgunyah Conservation Park, and the beach south-east of the Dog Fence (Caton et al., 2007). Recent surveys have indicated the presence of up to 10 pairs of hooded plovers between Twin Rocks and Granites (A. Loisier, AW NRM Board, pers. comm., 2010). Figure 5A and 5B show the location of Hooded Plover sightings at various beach locations in the AW NRM region, during surveys in August and December 2008, April 2009 and February 2010. According to counts during the past 3 years to 2010, numbers have been stable since 2008, and similar numbers of birds were recorded in the same locations during the three years (Table 3; A. Loisier, AW NRM Board, pers. comm., 2010). Table 3: Number of Hooded plovers counted during different seasons along the Yatala coast (between Granites and Head of Bight), from August 2008 to April 2010. Data from AW NRM Board. Survey Date Number of Hooded Plovers August 2008 13 December 2008 15 (including 1 juvenile) April 2009 14 March – April 2010 17 35 t t t Jerry's Beach (Bobs Kitchen) 1 t Jaxson's 2 2 2 2 2 26 2 3 2 2 2 Geues Hole tt t t Alinytjara Wilurara Natural Resource Management Shorebird Surveys Hooded Plover - Geues/Bob's Kitchen Month Transect Areas Granites - Coymbra August, 2008 Hilton - Twin Rocks December, 2008 ± Geues - Bob's Kitchen April, 2009 t 0 0.3 DISCLAIMER:The Alinytjara Wilurara Natural Resources Management Board, its employees and servants expressly disclaim all liability of responsibility to any person using the information or advice contained herein. 0.6 © Government of South Australia, through the Alinytjara Wilurara Natural Resources Management Board. This work is copyright. Apart from any use permitted under the copyright Act 1968 (Cwlth), no part may be reproduced by any process without prior written permission obtained from Alinytjara Wilurara Natural Resources Management Board. Request and enquires concerning reproduction and rights should be directed to the General Manager, Alinytjara Wilurara Natural Resources Management Board Kilometres Campsites February 2010 AWNRM_Boundary Extent of shorebird surveys t Map Production: Knowledge and Information, Alinytjara Wilurara Natural Resources Management Board Data Source: Topographic Data – From various State government departments Map Datum: GDA 94, Zone 52 Date: 15/6/2010 AW NRM Map 2010145_001 Tjitji Tjutaku Coombra t 22 2 t 3 1 2 1 2 1 1 2 t Alinytjara Wilurara Natural Resource Management Shorebird Surveys Hooded Plover - Granites/Coymbra Month Geues - Bob's Kitchen Granites - Coymbra August, 2008 Hilton - Twin Rocks December, 2008 February 2010 ± Transect Areas April, 2009 t 0 0.6 Kilometres Campsites AWNRM_Boundary Extent of shorebird surveys t 32 2 Granites 1 2 t Map Production: Knowledge and Information, Alinytjara Wilurara Natural Resources Management Board Data Source: Topographic Data – From various State government departments Map D atum: GDA 94, Z one 52 Date: 16/6/2010 DISC LAIMER:The Alinytjara Wilurara Natural Resources Management Board, its employees and servants expressly disclaim all liability of responsibility to any person using the information or advice contained herein. 1.2 © Government of South Australia, through the Alinytjara Wilurara Natural Resources Management Board. T his work is copyright. Apart from any use permitted under the copyright Act 1968 (Cwlth), no part may be reproduced by any process without prior written permission obtained from Alinytjara Wilurara Natural Resources Management Board. Request and enquires concerning reproduction and rights should be directed to the General Manager, Alinytjara Wilurara Natural Resources Management Board AW NRM Map 2010145_002 Figure 5: Location of Hooded Plover sightings during surveys at beaches in the AW NRM region, from August 2008 to February 2010. Maps provided by AW NRM Board. 36 Threats to populations of the ‗forementioned species are discussed in section 3.1.3 of Part 3. Other species of conservation concern recorded in the AW NRM region include Whimbrel, Great Knot, Ruddy Turnstone, Common Sandpiper, Pied Oystercatcher and Sooty Oystercatcher, all of which are listed as Rare under S.A. legislation), and all of which have been recorded recently in the Coymbra area (Dennis, 2008). Migratory albatrosses occasionally use the area but none were recorded in the 2008 AW NRM bird survey (Dennis, 2008). Species likely to pass through the AW NRM marine region (e.g. travelling, and/or feeding) include the Shy Albatross, Atlantic Yellow-nosed Albatross, Blackbrowed Albatross, Wandering Albatross, Sooty Albatross, Southern Royal Albatross and the Southern Giant-Petrel (Director of National Parks, Commonwealth DEH, 2005; AG DEWHA, 2009). 37 Part 2 USES OF THE AW NRM REGION – MARINE COMPONENT 2. Uses and Activities Recreational Fishing In both eastern Great Australian Bight and at the Head of the Bight, commonly caught species include mulloway, pink snapper, Australian salmon and sharks (Yalata Land Management, 2003; Wilson, 2007; The Fishing Guide, 2009). Remote ocean beach / surf fishing, particularly for mulloway, is widely promoted locally, interstate and internationally, and many fishers visit the area during the summer months to catch mulloway during their ―runs‖ along the shallow subtidal waters off surf beaches, and in nearshore reef ―gutters‖. Mulloway are fished every year in the AW NRM region. During the 2009-10 fishing season, a survey showed that 324 permits were issued, and 129 fishers (40% of total) were interviewed. Fishers stay, on average, 2 weeks at a time in the area and fish every day; some also fish at night. The 129 fishers who were interviewed during the 20091- season, fished a total of 621 hours over the sampled season. The survey did not include the busiest fishing period of the year (January holiday period), hence the total number of hours fished by all fishers over the whole season would be higher than that estimated from the survey. During the 2009-10 survey period, 81 legal-sized mulloway were caught from 7 locations in the AW NRM region (A. Loisier, AW NRM Board, pers. comm., 2010). Some fishing guides report that fishers expect to catch one or two ―good sized‖ (i.e. 1m) mulloway per day as the fish pass through the gutters. Some fishers return to the area every year, specifically to target mulloway in the GAB. On-line fishing forums and fishing reports (e.g. Fishnet; Got One; The Fishing Guide) and tourism materials for the Ceduna and Fowlers Bay area all promote the capture of mulloway in the Great Australian Bight. In this area, large (and some very large – to 30+ kg) mulloway are taken by individual recreational fishers, and charter boat fishers (e.g. Murton, 2003; Wilson, 2007; The Fishing Guide, 2009). Some are kept, and others released. In some years, there is a mulloway-catching competition run from October to January, by a hotel at Nundroo (Eyre Highway, 160 kilometres west of Ceduna). Issues associated with the capture of mulloway are discussed in section 3.1.1 Fishing in Part 3. Apart from the main targets (mulloway; large west Australian salmon and sharks such as gummy and bronze whaler), other species taken include pink snapper (including occasional very large individuals); mullet (often used for bait, along with small west Australian salmon), stingrays, and Port Jackson sharks. smooth hammerhead, school shark, and other shark species are also taken occasionally, but figures are not available for this report. The chimaera elephant fish / elephant shark (Appendix 6) is also taken seasonally in low numbers. 38 A number of recognised recreational fishing areas occur in the AW NRM region, including:  Head of the Bight (which is less accessible than some of the other areas to the east);  Dog Fence Beach (popular for fishing large mulloway – reportedly including specimens to 35kg – and large west Australian salmon;  Geues Hole, Jaxsons, Tjitji Tjutaku and other beaches off the Yalata Indigenous Protected Area (where a permit is required for fishing, and where fishers take mulloway, west Australian salmon, sharks, stingrays and reef fishes, including juvenile Western Blue Groper);  Coymbra (Coombra) Beach (also in the Yalata IPA), where surf fishers and anglers fish the deep lagoons, ―holes‖ and ―gutters‖ of the reef seaward of the beach. Examples of locations fished adjacent to the AW NRM area include the following (from Yalata Land Management, 2003; Hunt, 2006; Wilson, 2007; The Fishing Guide, 2009, and recreational fishing reports):  surf beaches around Point Fowler, Fowlers Bay and surrounds (e.g. Scotts Beach / Scotts Bay, Mexican Hat and Wandilla, popular fishing spots to catch west Australian salmon);  Fowlers Bay jetty (where Australian Herring / Tommy Ruff, whiting and mullet are caught, amongst other species);  Tuckamore Beach (where samson fish, west Australian salmon, mulloway and pink snapper area caught), Clare Bay and Cabbots Beach (where various reef fishes are taken from the ledges offshore); Charter boats visit a number of locations, particularly east of the boundary of the AW NRM region, in the Fowlers Bay and Nuyts Reef area, where, for species such as pink snapper, bag limits of small and large fish are easily caught. East of AW NRM region, offshore from Fowlers Bay, fishers catch reef species such as samson fish, Bight redfish, pink snapper, southern blue morwong, swallowtail, harlequin fish and sergeant baker (Hunt, 2006). Large whiting, snook and sand flathead are also caught in the area. southern bluefin tuna are targeted during the summer months (e.g. Hunt, 2006). Table 4 below shows the numbers of fishes caught (including numbers retained and released) in Fishing Block 1 during a National Recreational and Indigenous Fishing Survey (NRIFS), from March 2000 to April 2001. Fishing Block 1 constitutes State waters of the GAB, from Fowlers Bay in the east to the W.A. border in the west. In terms of numbers caught, the top species recorded during the survey were west Australian salmon, ―unspecified mullet‖ (including yelloweye mullet, jumping mullet, and/or sea mullet) and mulloway. Other species caught by recreational fishers in Fishing Block 1, but for which totals were not collated, included sea sweep, and flathead species (NRIFS data, 2001). According to Jones and Doonan (2005), no King George whiting, pink snapper, sea garfish, snook, southern calamari, blue swimmer crabs or sand crabs were caught in Fishing Block 1, but some of these species were caught in significant numbers in more sheltered waters to the east, such as the Ceduna and Streaky Bay area. 39 Table 4: Numbers of fishes caught, retained and released in Fishing Block 1 (Great Australian Bight) during the National Recreational and Indigenous Fishing Survey, 2000-2001 (adapted from Jones and Doonan, 2005). Species No. Retained No. Released Total No. Caught % released west Australian salmon 15,852 19,535 35,387 55.2 Australian herring (tommy ruff) 567 0 567 0 yellow-eye mullet 0 529 529 100 unspecified mullet (including yellow-eye mullet, jumping mullet, and/or sea mullet) 1,213 532 1,745 30.5 mulloway 898 2,033 2,931 69.4 whaler sharks (bronze whaler; dusky shark / black whaler) 53 0 53 0 rock lobster 112 0 112 0 Table 5 below shows the target and non-target effort used to catch fish species listed in Table 4. Target fishing effort is that effort expended by a recreational fisher, when a species is nominated as targeted, and caught during a fishing event (Jones and Doonan, 2005). The catch includes both the retained and released component. Non-target species fishing effort is that effort expended by a recreational fisher, when a species is caught and retained, but is not reported to have been targeted. Non-target effort does not include that effort when a species is released. Total effort is the sum of target and non-target effort (Jones and Doonan, 2005). In Fishing Block 1 at the Head of the Great Australian Bight, it was estimated that around 910 fishers fished a total of 1,815 fishing days during the 2000/01 survey period (Jones, 2009). In Fishing Block 1, fishers expended most effort (almost 23,000 hours fished) on west Australian salmon, over the period March 2000 to April 2001. An estimated total effort of 7,281 hours was spent fishing for mulloway, about 70% of this was targeted effort, during which fishers specifically sought to catch mulloway. ―Unspecified mullet‖ (more than one species) and tommy ruff (Australian herring) were caught but not targeted during the survey period. According to the results of the survey, all Whaler sharks were caught without being targeted specifically. Note that effort is 0 hours fishing for Yellow-eye Mullet (Table 3), because all individuals captured were reported to have been released. Of the species that were caught in abundance in the GAB during the NRIFS, west Australian salmon is a widespread, abundant migratory species in South Australia, and the numbers caught by recreational fishers in the GAB apparently from a small proportion (e.g. 3% in the NRIFS) of the State total. Similarly, the catches of the various mullet species appear to be insignificant on a State-wide scale (e.g. 0.4% of the annual State-wide catch of unspecific mullet, and 0.07% of the yellow-eye mullet catch, according to NRIFS survey results). 40 According to the results of the NRIFS, the recreational catches of mulloway in the GAB may be as low as 3% of the State total (i.e. 92,870 individuals were recorded from the entire State during the survey period, including released specimens, and 2,931 of these were caught in the Great Australian Bight). A more recent recreational fishing survey (Jones, 2009) reported an even lower proportion of the State catch of mulloway from the GAB during 2007/08, with 2% recorded for the entire ―West Coast‖ (includes waters east of the Fowlers Bay). During that survey period, the estimated total number of recreational fishers fishing in the Head of GAB area was 1,128 persons, who fished a total of 5,223 days during the 2007/08 survey period (Jones, 2009). The relatively low proportion of mulloway caught from the AW NRM region compared with other parts of South Australia may reflect the lower number of available records, and perhaps lower response rates from recreational survey participants in remote regions such as the GAB. Recreational catches of Whaler sharks (see Table 4) in State-managed waters of the Great Australian Bight may be as low as 1.5% of the State total, according to results of the NRIFS. Issues associated with the capture of Whaler sharks are discussed in 3.1.1 Fishing. A number of species that are caught by recreational fishers in the GAB, have not been listed in the results of the National Recreational and Indigenous Fishing Survey (Jones and Doonan, 2005) or the more recent 2007/08 survey (Jones, 2009) . Table 5: Target and non-target fishing effort (no. hours fished) for species caught in Fishing Block 1 (Great Australian Bight) during the National Recreational and Indigenous Fishing Survey, 20002001 (adapted from Jones and Doonan, 2005). Species Target Effort Non-Target Effort (hours fished) (hours fished) west Australian salmon 12,503 10,474 Australian herring (tommy ruff) 0 5,261 yellow-eye mullet 0 0 unspecified mullet (including yellow-eye mullet, jumping mullet, and/or sea mullet) 0 2,507 mulloway 5,057 2,224 whaler Sharks (bronze whaler and dusky shark / black whaler) 0 399 rock lobster 1,598 0 Charter boat fishing is not discussed in detail here because catch and effort statistics for that sector (e.g. Knight et al., 2007) list the GAB area under a category called ―Other‖, which includes the Head of the Bight, but also deeper waters off the western and southern sides of Eyre Peninsula. Fishers in that broad region, part of which includes the GAB, participate in fishing categorised as ―game fishing‖, ―inshore scale-fishing‖, ―offshore scale-fishing‖ and ―deepwater scale-fishing‖ (Knight et al., 2007). From September 2005 to June 2006, a reported 432 fishes (total weight 1.23 tonnes) were caught by charter boats in the ―Other‖ area, and from July 2006 to June 2007, the reported number was 614 fishes (6.3 tonnes) (Knight et al., 2007). According to Knight et al. (2007), pink snapper, Australian herring and King George whiting 41 were the only major commercial species taken in the ―Other‖ region during the survey periods (2004 – 2007), and this contrasts with the result of the NRIFS (see above), in which no King George whiting or snapper were reported to have been caught by recreational fishers in Head of the GAB (Fishing Block 1) during the survey period 2000-01 (with results from 2000/01 being considered typical of annual catches). This indicates that the majority of the charter boat catch from the region classified as ―Other‖ would come from waters away from the Head of the Bight. Commercial Fishing The State waters component of the Great Australian Bight is not a major fishing area, possibly due to the rough conditions, lack of easy access, remoteness from ports, and presence of GAB Marine Park (see Part 3). The AW NRM marine region covers three of the 1-degree fishing blocks (blocks 1, 2 and 3) of the multi-species, multi-method South Australian Marine Scalefish Fishery (Figure 6). Much of the species-specific data is confidential, because less than 4 fishers fish for those species in each block (SARDI Aquatic Sciences statistics section, pers. comm., 2009). During the mid-1990s, the major species taken were school shark and gummy shark (e.g. approximately 10 - 20 tonnes of each species taken per annum from each block in some years) and whaler sharks (i.e. bronze whaler and/or black whaler – several tonnes per annum taken from blocks 2 and 3). It is noted that an Offshore Constitutional Settlement (OCS) during the late 1990s between the Commonwealth Government and South Australia, resulted in management of commercial shark fishing in South Australia becoming a Commonwealth responsibility, and the agreement was signed in 2000. Under the OCS, the Commonwealth was given jurisdiction from the low water mark to the boundary of the Australian Fishing Zone, for commercial fishing of sharks. The Department of Agriculture, Fisheries and Forestry conducted a buy-out of 40 operators from the fishery. Therefore, catch and effort data for commercial capture of school sharks and Gummy sharks in South Australian waters from that time onwards (and including the 2000s) were collated by the Commonwealth, and block-specific data within the Great Australian Bight region are not available. The shark fishery is discussed in more detail in the paragraph below. Other sharks are taken in low numbers in the South Australian Marine Scalefish fishery, and low quantities of scalefish such as Australian Salmon and mulloway are also taken. Most scalefish data from 1997/98 to 2007/08 are confidential (SARDI Aquatic Sciences catch and effort statistics, 2009). In both State-managed waters (see above) and Commonwealth-managed waters of South Australia, sharks are targeted in the Shark Gillnet and Shark Hook Sectors (formerly called the Southern Shark Fishery) of the Commonwealth-managed Southern and Eastern Scalefish and Shark Fisheries (SESSF). In deeper waters of the Great Australian Bight (i.e. not in AW NRM waters) , the Great Australian Bight Trawl Sector (GABTS) of the SESSF operates in Commonwealth-managed waters. The Australian endemic gummy shark is the main target species in the Shark Gillnet and Shark Hook Sectors, and represented approximately 77% - 80% by weight of the total shark catch in the fishery in 2007 and 2008 (McLoughlin and Wood, 2009). School sharks are taken as byproduct in the fishery, and have comprised approximately 8% - 10% of the total catch by weight in recent years (McLoughlin and Wood, 2009). Both of these sharks are taken in waters of the AW NRM region, at the Head of the Bight, as mentioned above, but in low numbers. The main fishing areas for sharks are western and south-western Eyre Peninsula, southern Kangaroo 42 Island; waters offshore from the Coorong, and the upper and mid South East of South Australia (McLoughlin and Wood, 2009). The fishery is also active in Bass Strait (particularly the eastern side), and eastern Tasmania. Issues associated with the fishing of school shark in southern Australia are discussed below, in section 3.1.1 Fishing in Part 3. N Head of Bight Nullarbor 01 02 03 04 05 06 12 0 100 Fowlers Bay 07 13 200 Kilometers Figure 6: South Australian Marine Scalefish Fishery 1-degree fishing blocks, showing the 3 blocks (01, 02, 03) included in the AW NRM region. A list of species taken commercially in South Australian Marine Scalefish Fishery blocks 1,2 and/or 3 is provided below (Table 6). Other than school sharks and gummy sharks, most of these species have been taken in low (some very low) quantities during the past decade. The population of mulloway at the Head of the Bight was reported to have been targetted heavily by commercial fishers during the 1990s, with initially high catches and catch rates reported, prior to reported depletion of the stock (K. Evans, SARDI, pers. comm., 1996). Catch numbers of mulloway taken per annum are not available for this report, because data from 1997/98 to 2008/09 are confidential (SARDI Aquatic Sciences catch and effort statistics, 2009), other than a 1.5t catch of mulloway in 2000/01, from the region which includes blocks 1,2, and 3 combined. 43 Table 6: Bony and cartilaginous fishes and invertebrates taken by the commercial Marine Scalefish fishery, in fishing blocks 1,2 and 3 (SARDI Aquatic Sciences data, 1997 – 2008) Bony Fishes Cartilaginous Fishes Invertebrates west Australian salmon school shark cuttlefish mulloway gummy shark southern calamari yellow-eye Mullet whaler sharks (bronze whaler &/or black whaler) sand crab Garfish ―other sharks‖ King George whiting ―rays and skates‖ snapper Snook Sweep wrasses (―parrotfish‖) leatherjackets ―other / mixed species‖ Catches of whaler sharks in South Australian waters are usually reported as bronze whaler, but the catches also contain black whaler C. obscurus (Jones, 2008). There are 8 fisheries in South Australia (State and Commonwealth combined) that catch whaler sharks, plus 2 other fisheries in which they are bycatch, plus incidental mortality from aquaculture operations. Examples of catches include Marine Scalefish Fishery long-line fishery in South Australia; GAB Trawl Fishery (bronze whaler is a by-product from fishing in shelf waters); Gillnet Hook and Trap Fishery in Commonwealth waters; and bycatch in the southern bluefin tuna fishery (including attraction to towed cages in and around the GAB). In the State and Commonwealth fisheries, tighter controls in 2002/03 on school shark and gummy shark fishing resulted in increased targetting of whaler sharks (Jones, 2008). Rock lobster are taken in low quantities in the Head of Bight area, as part of Region A of the Northern Zone of the South Australian Rock Lobster Fishery. Figure 7 below shows the geographic extent of Region A, of which the marine region of AW NRM forms a small part. Catches for fishing blocks 1, 2 and 3 – from east of the Head of Bight through to the S.A. / W.A. border – are not available, but it is noted that less than 10% of the entire Northern Zone catch is taken from Region A (e.g. 4% in 2006, and 8% in 2007: Linnane et al., 2008, 2009) and these three blocks form only a small portion of Region A (Figure 7). The Region A catch was higher during the mid 1970s (e.g. almost 150t in 1975) and mid 1990s (e.g. 172t in 1993), as were catch rates (e.g. > 2kg per pot lift in 1990, 1992 and 1993), but catches (especially) and catch rates are now lower, compared with those from other parts of the Northern Zone. In 2006 and 2007, the total Region A catches were 20 tonnes and 36 tonnes respectively, from all 16 fishing blocks in Region A combined (Linnane et al., 2008, 2009). 44 N Head of Bight Nullarbor 01 02 Fowlers Bay 03 Ceduna 09 04 61 05 06 07 62 12 13 24 3 0 200 08 14 25 3 400 Kilometers Figure 7: Region A (numbered fishing blocks) of the South Australian Rock Lobster Fishery, showing the 3 blocks (01, 02, 03) included in the AW NRM region. Abalone are taken in low quantities in the Head of the GAB, as part of Region B of the Western Zone of the South Australian Abalone Fishery. Region B is a broad area that encompasses waters from the S.A. / W.A. border, eastwards to the Point Brown area near Smoky Bay. As part of that broad area, the marine component of the AW NRM region lies within the Head of Bight and Nullarbor fishing area known as Abalone Fishing Area 1A, nested within Abalone Fishing Area 1, which runs from the S.A. / W.A. border eastwards to Point Bell (Figure 8). Catch and effort data specific to Abalone Fishing Area 1A are not available for this report, but it is noted that catches of greenlip abalone from Fishing Area 1 as a whole have been stable at about 3t per year since the mid 1990s. Slightly higher catches (between 5t and 18t) of Greenlip were taken in some years of the 1980s and early 1990s (Chick et al., 2007). Catches of blacklip abalone are slightly higher than those of greenlip, and although less than 5t per annum have been taken from Fishing Area 1 since the mid 2000s, catches between 5t and 10t per annum were taken in most years from the early to late 1990s. About 6t of blacklip were taken in 2002 from Fishing Area 1, and annual catches have been lower since that time, to the limits of available data (2007). The blacklip catches from Fishing Area 1 have represented between 10% and 40% of the annual Region B Blacklip Abalone catch since 2000 (Chick et al., 2007). 45 N 1A 1D 2 1B 2 1E2 1C 2 0 100 200 Kilometers Figure 8: Fishing Area 1 (numbered fishing blocks) of Region B of the Western Zone of the South Australian Abalone Fishery, showing the fishing area (1A) of which the AW NRM region forms a part. The annual catch of both abalone species from Fishing Area 1 forms a relatively small proportion of the Region B quota of 41.4 tonnes of greenlip and blacklip abalone combined (Chick et al., 2007), and the catch from the AW NRM region (Fishing Area 1A) would form an even smaller proportion, being the most remote location in Fishing Area 1, and thus least accessible for commercial abalone fishers. From the early 1990s to the present, between 50% and 80% of the abalone catch in Region B has come from well east of the Bight, in Fishing Area 2. Giant crab (Pseudocarcinus gigas) is another invertebrate species taken in the Great Australian Bight, but a recent stock assessment report (Currie and Ward, 2009) indicates that the fishery operates in waters south-east of the Head of the Bight, and off western and southern Eyre Peninsula, in deeper (Commonwealth-managed) waters. This species is not taken in the marine waters of the AW NRM region. In deeper waters seaward of the AW NRM region, the central zone of a Commonwealthmanaged Great Australian Bight trawl fishery operates, which is part of the SESSF (see above). Deepwater flathead Neoplatycephalus conatus and Bight redfish Centroberyx gerrardi are the principal species taken (Lynch and Garvey, 2003; AFFA, 2004), and have been trawled sporadically in the GAB since the early 1900s (Kailola et al., 1993). A dedicated fishery was set up and managed as a developmental fishery in 1988, and since then, a permanent fishery has been established with steadily increasing catches of both species (Klaer, 2007). In shelf waters of the GAB, the trawlers have traditionally operated at about 120–160m depth, but most of the effort in the fishery is now concentrated on the upper continental shelf and slope, in depths from 100m to 400m (AFMA, 2008c), and in particular, the shelf break south of the Head of the Bight, in Commonwealth-managed waters (Morison et al., 2009). During spring and summer in recent years, integrated scientific monitoring program (ISMP) observer reports indicate that 9 of the 10 46 licensed vessels consistently work in the GAB trawl fishery, with most vessels targeting deepwater flathead and Bight redfish on the outer shelf (100m – 220m deep). Some vessels target blue grenadier, gemfish and ling. Vessels work across the whole of the GAB from 125o E to 134o E, but the effort expended to catch deepwater flathead, Bight redfish and ocean leatherjacket is concentrated at the shelf break of the central GAB (Lynch and Garvey, 2003), including Commonwealth-managed waters seaward of the AW NRM region. The shelf trawl fishery in the GAB operates year-round, and deepwater flathead catches and catch rates peak in October–December and those for Bight redfish peak in February–April (AFFA, 2004). There appears to be an inverse relationship between redfish and flathead catchrates that is probably environmentally driven (AFFA, 2004), and may relate to variable patterns over time in dispersal or aggregation for feeding, in relation to variable productivity. In some years (e.g. 2002), significant tonnages of other species are taken in the central zone of the fishery (e.g. Lynch and Garvey, 2003). Other important continental shelf species taken by the GAB trawl fishery include ornate angel shark (Squatina tergocellata), ocean leatherjacket (Nelusetta ayraudi), brown-spotted (yellow-spotted) boarfish (Paristiopterus gallipavo), jackass morwong (Nemadactylus macropterus), western gemfish (Rexea solandri). Other species taken in lower quantities in the shelf part of the fishery include Gould‘s squid (Nototodarus gouldi), knifejaw (Oplegnathus woodwardi), latchet (Pterygotrigla polyommata), sawsharks (Pristiophorus species), whiskery shark (Furgaleus macki) gummy shark (Mustelus antarcticus), and bronze whaler Carcharhinus brachyurus (Lynch and Garvey, 2003, and fisheries bycatch data). Apparently, several tonnes of bronze whaler per year are taken by the GAB trawl fishing in the shelf waters of the GAB (Figure 3.11 in Jones, 2008). In recent years, gurnard perches (Neosebates species), ocean perch (Helicolenus percoides), long-finned / black-spot boarfish (Zanclistius elevatus) and queen snapper / southern blue morwong (Nemadactylus valenciennesi) have been increasingly retained when caught (ISMP data, 200506). Pelagic schooling species such as redbait and jack mackerel are also regularly caught. Annual landings of the shelf species have varied, largely depending upon the amount of fishing effort on the continental shelf (AFFA, 2004). There is a large discarded bycatch of latchet, wide stingaree (Urolophus expansus), spikey dogfish (Squalus megalops), juveniles of ocean leatherjacket (Nelusetta ayraudi) and many other species, as well as attached benthos (see Part 3). Other fisheries which operate in the Great Australian Bight include the Gillnet Hook and Trap sector of the Southern and Eastern Scalefish and Shark Fisheries); Small Pelagic Fishery Zone B – only operators who also hold concessions in the Great Australian Bight Trawl Fishery (part of the Southern and Eastern Scalefish and Shark Fishery); Southern Bluefin Tuna Fishery; the Southern and Western Tuna and Billfish Fishery; the Skipjack Tuna Fishery, and the Southern Squid Jig Fishery (Director of National Parks, 2005), but these fisheries do not operate in the shallow waters of the AW NRM region. Approximately 99% of the annual catch of southern bluefin tuna in southern Australia comes from the continental shelf break area south-east of the Head of the Bight (and west of Anxious Bay), in Commonwealth-managed waters (Phillips et al., in Wilson et al., 2009). In some years, this species is also caught in deeper waters south of the head of the Bight (e.g. Figure 12 in Ward et al., 2006). Mid-water trawls have also begun to target comparatively small pelagic species such as jack mackerel (Trachurus declivis) and blue mackerel (Scomber australasicus) over a broad area in GAB waters, and this activity increased during the 2000s (AFFA, 2004; Morison et al., 2009). 47 Coastal and Marine Recreation and Tourism The migration of Southern Right Whales to the Head of the Bight between May and October each year is of significant tourism value. Whale watching / site seeing in Yalata lands and Nullarbor National Park is one of the primary recreational activities in the area (Caton et al., 2007). The near-shore whale watching opportunities afforded by the cliffs and viewing platform in the area, make the GAB of State, national and international significance for whale watching in a non-invasive way that does not disturb the natural behaviour of the whales. According to Yalata Land Management (2003) and O‘Connor et al. (2004), over 15,000 visitors come to the area during the whale-watching season. The Head of the Bight is part of South Australia‘s Southern Right Whale Trail, which stretches from Robe in the south-east of the Great Australian Bight in the west, and interpretative signs and information are provided at whale-watching sites along the route. The interpretative centre at the Head of Bight is operated by Maralinga Tjarutja Inc., and contains information about the social and natural history of the area. A number of display boards discuss the habits, behaviours and other scientific curiosities relating to the southern right whale. Boardwalks lead from the centre to the Head of Bight where the whales can be observed. White sharks, Australian sea lions and other fauna are also seen occasionally from the Head of the Bight, and this is promoted by Tourism South Australia as part of the whale-watching experience. During the whale watching season, an aviation company at Nullarbor Roadhouse offer 20 to 30 minute flights over the area. Visitors are still permitted to enter the Head of Bight during the summer months (when the whales are not present). In addition to whale watching, camping is also one of the coastal uses in the area, and is often associated with surf fishing (e.g. Wilson, 2007; The Fishing Guide, 2009). Fishing and camping in Yalata lands are popular activities (Caton et al., 2007). The Yalata Indigenous Protected Area is reported to provide some of the best natural dune camping areas available in South Australia, and the camping areas are adjacent to renowned remote surf fishing spots. Fishers use 7 main camp sites (Hilton, Bob‘s Kitchen, Jaxon‘s, Geues, Tjitji Tjutaku, Coombra and Granites) spread along the beach, and some take four-wheel drive motorbikes (quad bikes) to travel from the camp sites to the beach, and to move along the beach. Camping is permissible in designated camp sites, and a permit is required. Within the Yalata Indigenous Protected Area, vehicles are permitted to enter the beach at locations named Granites, Bob's Kitchen and Hilton, and other campsites (Yalata Land Management, 2003). Apart from the popularity of whale-watching, tourism materials promote the coast of the Head of Bight and Nullarbor for the spectacular scenery. The coastline in this region is nationally and internationally renowned for its scenic amenity and wilderness value. In the east, the very large sand dune formations, and in the west, the 200km of towering cliffs facing into the Southern Ocean, are promoted for sightseeing and ecotourism experiences. The vast dry and wet cave systems in the limestone karst regions below the Nullarbor are popular with local and international divers as well as tourists, but it is noted that some caves have been closed in recent years due to safety risks. Adjacent to the Nullarbor National Park and due south of Yalata Roadhouse, the Great Australian Bight coast is promoted for ―excellent wilderness fishing‖ (Eyre Peninsula Tourism Association, 2000). To the west, the scenic coastal look-outs of the Bunda Cliffs, as well as the underground caves, and blowholes of the Nullarbor National Park, are promoted for ecotourism. 48 Aboriginal Heritage The Aboriginal history and cultural value of the Head of the Bight, which includes the Wirangu in the east and the Mirning in the west, is discussed in other reports (including Cane, 1989, cited by Edyvane and Andrews, 1998 and Caton et al., 2007) and will not be reiterated here. It is noted, however, that the coastal area has long been important to both groups, particularly during the spring and summer, with coastal resources including seals / sea lions, shellfish, bony fish and birds (e.g. Australian Bustard). In the past, when much of the coastline was inaccessible because of the cliffs, occupation focused around fresh water soaks at the Head of Bight, Eucla and Merdayerrah Sandpatch. Rockholes, a significant karst feature at places along the coast and further inland, also have both natural and cultural value, being the main places where water can collect on the dry surface. The caves of the Nullarbor, which provide habitat for unusual and endemic fauna, are considered to be very important cultural heritage for the indigenous Mirning people. At the Head of the Bight, the Yalata swamp and dune system is of cultural significance to Anangu and Wirangu people and other Aboriginal communities. This area contains the traditional waterhole and meeting place called Illcumba. The area of significance to Anangu people runs north and NE from the swamp, up to and beyond The Lands‘ northern boundary (Yalata Land Management, 2003). Caton et al.‘s (2007) Far West Coastal Action Plan mapped the parts of the GAB coast that are Indigenous Protected Areas, owned by Aboriginal Lands Trust, and leased to Yalata Community Incorporated. The Yalata people on the Yalata Aboriginal Lands inhabit the Head of the Bight region. According to Yalata Land Management (2003), the core population is approximately 250 people, and they maintain an active cultural life. The Yalata people regard themselves as the southern Anangu and speak a southern dialect of Pitjantjatjara as their first language (Yalata Land Management, 2003). The main population centre is around Yalata township, and Yalata Roadhouse, but there are extensive track systems to the north-east and north-west. The community hunts within all of The Lands depending on seasonal availability. South-west from Yalata township, the Yalata Beach area, is a usual recreation site for community members (Yalata Land Management, 2003). The Mirning, Wirungu, Maralinga Tjaruta and Anangu hold indigenous cultural interests in the waters of the Bight. For the Anangu based at Yalata, interests focus mainly on the inshore reef areas within State waters (Jeremy LeBois, Yalata Community, pers. comm., cited by Director of National Parks, Australian Government DEH, 2005). Sites of particular significance for the Aboriginal community have also been identified through literature reviews and consultation, and a number of actions have been recommended to ensure that such significant sites (and any others that Aboriginal communities should wish to register with the Aboriginal Affairs and Reconciliation Division) are better identified, managed and protected (Caton et al., 2007). In addition to the Yalata IPA, some other examples of significant indigenous heritage sites, according to Caton et al. (2007) include the Wahgunyah Sandhills, and the coastal area southeast of Dog Fence beach. 49 Research In 2008, a coastal field study was undertaken to determine the current status and distribution of the threatened coastal sea birds white-bellied sea eagle, osprey and peregrine falcon in the AW NRM region, and to identify habitat threats (Dennis, 2008). The field study also aimed to identify locations where little penguins may be aggregating ashore, and to assess breeding status where possible. Counts of other coastal birds and sea birds were also made. The conservation organisation Birds Australia also undertakes shorebird counts on an irregular basis, at two sites in the AW NRM area: one at the very head of the Bight, and the other on the western side of the Head of the Bight where the coastal direction turns from south-west to westwards. Recent counts of coastal birds, included the hooded plover, have been undertaken in the AW NRM region (A. Loisier, pers. comm., 2010). Other coastal research projects will not be discussed here. A number of marine research projects have been carried out in the AW NRM region during the past two decades. Researchers have been undertaking a detailed and long-term study of the southern right whales at the Head of Bight since 1991. This research includes photographing individuals for a photo catalogue (to record the callosities, unique markings which can be used to assist identification over space and time), recording behaviour, mapping distribution, and assessing how the whales use the area over years and decades. As at 2006, the photo archive comprised more than 20,000 images, spanning more than 15 years. In 2006, there were 683 individual right whales in the photo catalogue, and gender was known for ~60% of these. Details such as callosities, dorsal and ventral pigmentation, ventral anatomy and scarring are recorded for each image. The re-sighting history (date, location, status) is maintained in a database, and a computer-assisted photo matching system has been developed (Burnell and Pirzl, 2006). Through this photo identification project, researchers have been able to determine how often the whales visit the Great Australian Bight Marine Park. During the annual surveys, aspects of Southern Right Whale movements, distribution and behaviour are recorded, and a range of environmental conditions are sampled. The photo archive and associated data sets (life history, habitat use, census, and behavioural data sets) has enabled the study of coastal residence periods, visitation rates and habitat preference, population size, behaviour and social interactions, migration / long-distance movements and travel speeds, and various aspects of the population dynamics, such as growth rates, breeding age, reproductive rates, frequency of calving and age at weaning (Burnell and Pirzl, 2006; research by Burnell and Pirzl, cited in Department for Environment and Heritage, 2006). During the early to mid 2000s, the Head of Bight was one of the three main areas in southern Australia in which research on the spatial ecology of southern right whales was undertaken. The surveyed area comprised waters from the shore to 3 km seaward along 15km of coastline, which encompassed almost the entire calving ground defined from previous aerial observations and shore-based study (Bannister, 1979; Ling and Needham, 1985; Burnell, 1999, all cited by Pirzl, 2008). During the project, a species-habitat model was developed, and habitat selection was investigated, including extent of calving grounds, spatial arrangement within calving ground aggregations, spatial extent of movement range of adults, long-term patterns of spatial use within range, and factors influencing the selection of locations as calving grounds. Techniques used in the study included point pattern analysis, spatially sensitive generalised linear modelling, univariate statistics, nonparametric analysisof variance and rules-based environmental envelope models (Pirzl, 2008). 50 In additon to the habitat usage study, a satellite telemetry project was also underway during the mid 2000s, so that researchers could examine connections with offshore habitat components as an extension of coastal habitat preference studies across southern Australia (Pirzl, 2004; Pirzl and Burnell, 2004). A project has also been undertaken to determine whether genetic differences exist between stocks of southern right whales aggregating at different locations along the southern Australian coastline (research by N. Patenaude and R. Harcourt, Marine Mammal Research Group, Macquarie University, 2001-04; Patenaude and Harcourt, 2006). At a broader scale, aerial surveys of southern right whales throughout south western Australia and western part of south Australia have been ongoing since 1976, and this long term project includes population estimates and maintenance of a photo-identification catalogue (research by J. Bannister, West Australian Museum, cited by DEWHA, 2010). During the 2000s, SARDI Aquatic Sciences tagged Australian sea lions with satellite tags, at a 80m high section of the Bunda Cliffs along the central GAB coastline. The project, for the Commonwealth and South Australian departments for Environment and Heritage, aims to understand more about the movements and foraging habitats of the sea lions, and what role the Marine Mammal Protection Zone of the GAB Marine Park plays in protection of sea lions (SARDI Aquatic Sciences, 2006). State Emergency Service vertical rescue experts and volunteers also assisted the project, which tagged female members of one of the seven colonies that breed along the 200 kilometre stretch of Bunda Cliffs. Within a fortnight of the tagging, one of the sea lions had already moved more than 180 km from the Bunda Cliffs base, while another had remained close to the colony (SARDI Aquatic Sciences, 2006). As part of the project, spatial analyses will be used to determine distance and direction of travel from colonies by males and females of different age classes, and depth of foraging (Goldsworthy et al., 2009a). Another survey, in 2008, aimed to opportunistically locate previously known and unknown Australian sea lion haul-out locations in the AW NRM region, and to record evidence of breeding activity and data on all age-classes present, where it was safe to do so (Dennis, 2008). The Bunda Cliffs colonies have not commonly been surveyed over the past two decades, due to inaccessibility and remoteness. For example, between 1985 and 2008, two of the Bunda Cliff colonies were sampled during two breeding seasons; four colonies sampled during three seasons, and one colony sampled four times (Goldsworthy et al., 2009a). Other research on Australian sea lions includes PhD projects on the degree of population subdivision / meta-population structure among breeding sites, including a comparison of the Bunda Cliffs populations with others in South Australia and Western Australia (research by University of Adelaide, SARDI Aquatic Sciences and Macquarie University, cited by Goldsworthy et al., 2009a). There is also an investigation by SARDI Aquatic Sciences, the South Australian Museum and WA Fisheries into dietary preferences of Australian sea lions across their range (Goldsworthy et al., 2007). In recent years, a recreational fishing survey has been undertaken by the AW NRM Board, in conjunction with SARDI Aquatic Sciences. The format is similar to that used in a Statewide recreational fishing survey (Jones, 2008). Information is being collected on mulloway, and other commonly fished species in the AW region. During the 2009-10 fishing season, 40% of those who fished in the AW NRM region was targeted. A promotional brochure about the survey has been distributed. Researchers are collecting information on number and species of fish caught, stage of maturity (where possible to assess), and earbones are being collected for ageing work, and comparisons with mulloway populations form other parts of South Australia. Data from the 51 first year are currently being analysed, and the survey will continue in future years (A. Loisier, AW NRM Board, pers. comm., 2010). Other studies in shallow shelf waters of the Head of the GAB include studies of the marine plants in the AW NRM region, including (i) phylogenetics and phylogeography of marine benthic macroalgae of the Great Australian Bight: a PhD study which is using molecular tools to provide a taxonomically updated survey of the marine flora of the head of the GAB and vicinity; testing the hypothesis that the Bunda Cliffs acts as a barrier to gene flow between western and eastern GAB populations, and testing whether species of macroalgae considered ―relicts‖ trapped at the northern reaches of the GAB and SA gulfs belong to ancient populations, or whether they arrived recently as ―long range dispersals‖ (University of Adelaide and S.A. Plant Biodiversity Centre project); and (ii) status and taxonomy of Sargassum species in the AW NRM region (e.g. Gurgel, 2009). In deeper waters seaward of the AW NRM region boundary, the GAB Industry Association, CSIRO, and external consultants are engaged in a number of projects, including fisheryindependent surveys of Bight redfish and deepwater flathead stocks; catch and bycatch monitoring programs, and characterisation of the seabed using multi-scale maps, to develop habitat classifications and correlate with catches. Also in deeper waters of the GAB, white sharks (CSIRO data) and shortfin mako sharks (SARDI Aquatic Sciences and Flinders University data) are being tracked, to learn more about seasonal and depth-related movements. Diet and DNA samples are also being taken from sharks in the GAB, and in other parts of South Australia (henrythesealion.com., 2010). In Western Australia, the multi-participant Marine Futures program (Government of Western Australia, 2006) has been mapping habitats and undertaking biodiversity surveys in the marine waters associated with 5 participating NRM regions. In a mid-2000s workshop for prioritising sites (Government of Western Australia, 2006), areas in the Eucla Bioregion (includes Baxter Cliffs, Twilight Cove and Eyre coast) were ranked as low priority for surveys, due to high cost and relatively low impacts, other than fishing. Ports, Shipping and Boating There are no major or minor ports in the marine waters of the AW NRM region. The nearest ports are Thevenard (near Ceduna) in the eastern GAB, and Esperance in W.A., in the west. Of the 9,185 Australian-owned commercial or trading ships registered in Australia at June 2004, 1,208 are located in Western Australia and 643 in South Australia (ABS data, cited in ABS, 2007). About half (311) of those vessels in South Australia are used for fishing; 284 for recreation, and only 47 are commercial or trading vessels. In Western Australian, 53% (or 642 vessels) are registered for recreational purposes; 34% used for fishing, and 12% are used for trading and other commercial purposes (ABS data to June 2004, cited in ABS, 2007). In South Australia in 2001, domestic sea freight transported 7 million tonnes of goods, over 10,184 million tonne-kilometres. The same quantity of goods was also transported domestically from Western Australia, over 33,691 million tonne-kilometres (ABS, 2007). When travelling through Australian coastal waters, commercial vessels report their position each day to the 52 Australian Maritime Safety Authority as part of the AUSREP ship reporting system. Using ship position data, it is possible to determine the major shipping routes of vessels moving through the Great Australian Bight in South Australia and Western Australia (Figure 9; Gardner et al., 2006). It is clear from the figure of shipping routes, that ships bypass the shallow waters of the Great Australian Bight. The busiest shipping routes run east–west through the Southern Ocean, in Commonwealth-managed waters. Major shipping routes stem from Gulf of St Vincent and Spencer Gulf in South Australia, and from Albany, Fremantle and Perth in Western Australia. Collectively, this shipping traffic includes international and national cargo trade, and some passenger services. Within the Great Australian Bight, the majority of the shipping activity occurs in the south-eastern area, between the Ceduna / Streaky Bay region to bottom of Eyre Peninsula. Figure 9: Shipping routes (determined from ship position data) in the Great Australian Bight. Data from Australian Maritime Safety Authority. Figure is from Gardner et al. (2006), based on data from the Australian Maritime Safety Authority. Mining, Minerals, Oil and Gas Exploration Currently there is no offshore production of oil and gas within the Great Australian Bight. However, oil and petroleum are imported through ports east (Port Lincoln) and west (Esperance) of the GAB (Australian Association of Port and Marine Authorities data, 2003, cited by Gardner et al., 2006). Although there is currently no offshore production within the GAB, exploration activity has been undertaken, including the release of exploration acreage and permits, seismic and magnetic surveys, and drilling of wells (Figure 10; DITR, 2003; Gardner et al., 2006; Fugro, 2008; Struckmeyer, 2009). The AW NRM region forms part of the Eucla Basin and the Bight Basin, and the Bight Basin has been identified as a ―frontier area‖ with petroleum 53 potential. This basin spans over 800,000 square kilometres, from the southern tip of Western Australia to the western edge of Kangaroo Island, and varies in depth from 200 metres to over 4000 metres (Geoscience Australia, 2010). From the late 1960s to the early 1990s, only nine offshore wells were drilled in the Bight Basin. However, since 2000 the Australian Government has focused on the Great Australian Bight for the release of exploration acreage, with three large exploration permits awarded in the Ceduna Sub-basin (Figure 11) and several smaller permits granted. The exploration programs associated with these permits include the acquisition and reprocessing of over 18,000 line kilometres of 2D seismic data, and geological and geophysical studies. Exploration interest focused largely on the Ceduna and Duntroon Subbasins (Geoscience Australia, 2010). In 2003, Woodside and partners drilled the unsuccessful Gnarlyknots 1 in the Ceduna Sub-basin, the first well in the basin since 1993. By November 2007, all exploration permits in the Bight Basin had been relinquished. In early 2007, Geoscience Australia undertook a 3 week sampling survey which identified potential source rocks in the basin. A total of 37 dredge samples were taken, resulting in 259 samples that underwent organic geochemical analyses. These analyses showed that a set of samples from the western Ceduna Sub-basin have characteristics of source rock, with high organic carbon contents and the potential to generate liquid hydrocarbons (Geoscience Australia, 2010). It is noted that some of the exporation activity in the GAB has taken place in the Commonwealthmanaged waters of the Great Australian Bight Marine Park (see section 3.1.8 in Part 3). Note that none of the exploration activities have taken place in the shallow waters of the AW NRM region. Petroleum exploration has been undertaken in deeper Commonwealth-managed waters. A 54 B C Figure 10 A,B,C: Exploration permits and exploration wells (A) and seismic surveys (B,C) in the Great Australian Bight. Figures A and B from Gardner et al. (2006), based on data from DITR, NIMA and Geoscience Australia; Figure C from Geoscience Australia (2010). 55 Figure 11: Location of the Bight Basin and sub-basins. Figure from Geoscience Australia (2010). 56 Part 3 POTENTIAL THREATS AND THREATENING PROCESSES IN THE AW NRM REGION – MARINE COMPONENT 3 Potential Threats Fishing Fishing in South Australian waters: A number of fish species of conservation concern are taken by commercial and/or recreational fishers in State-managed waters of the Great Australian Bight. Examples include mulloway, Western Blue Groper, Harlequin Fish, Southern Blue Morwong, school shark, Whaler Sharks and Smooth hammerhead. Although only 3% or less of the annual recreational catch of mulloway is estimated to be taken from the Great Australian Bight (see section 2.1.1 Recreational Fishing), and the commercial catch is also low as a proportion of the State total, any catch of this species should be closely monitored. Mulloway have vulnerable population characteristics, and are considered susceptible to over-exploitation at international, national, and Statewide scales. Ward and Butler (2006) described mulloway in southern Australia as a ―sensitive species that is highly targeted by recreational fishers‖. In other parts of the world, such as East Cape of South Africa, mulloway is described as being very highly exploited, with very high vulnerability to overexploitation, and poor stock status (Britz et al., 2001), and the situation is similar in South Australia (Ferguson et al., 2008; Baker, 2009). In Western Australia, mulloway is classified by the Department of Fisheries as a ―high risk‖ species. There are few parts of South Australia where this species occurs in significant numbers, other than the Murray Mouth area (in particular), southern Gulf St Vincent / Fleurieu Peninsula, north-eastern Spencer Gulf, and the South East. The mulloway in the Great Australian Bight may be a genetically isolated population (Jones, 1991), which may not mix with mulloway from other parts of the State. Observations from fishers in the Head of the Bight indicate a westward movement of fish in the spring to summer period and a reverse movement in the autumn to winter period (Jones, 1991). This species undertakes seasonal movements along the coast, possibly in response to the seasonal movements of their prey, such as pilchards (Gomon et al., 1994). Fishers report that mulloway movements are influenced by many factors, such as moon phases, tides and storms. Mulloway grow rapidly up to the end of their sixth year of life. After that time, growth slows markedly. About 80% of the maximum size is reached by 6 years of age, but the maximum age is about 32 years (Farmer et al., 2005; Farmer, 2008). In Western Australia, a research project by B. Farmer (Farmer, 2008) estimated that length at first maturity is typically 90.3cm for female mulloway and 88cm for males on the west coast of W.A. (equivalent to 5 years of age or more), but only 49.3cm for females and 41.9cm for males (3 years of age) at sampled locations on the southern coast of W.A., such as Oyster Harbour (Farmer, 2008). Spawning aggregations are not typical, but have been reported in some areas (e.g. Swan River estuary, in Western Australia: Farmer, 2008; Parsons et al., 2007, 2009). Year class strength is variable, and strong year classes, or their absence, may be related to freshwater inflow (Ferguson et al., 2008). 57 More genetic research is required to determine the status of the population in the Great Australian Bight, but it is noted that preliminary stock identification studies indicate that the genetic composition of mulloway along the south coast of Western Australia differs markedly from that of mulloway along the west coast of W.A., and that there is also a degree of genetic differentiation between two populations of mulloway along the western coast of W.A., although to a lesser degree when compared with the south coast (Farmer, 2008). Table 1 in section 2 above shows that 69% of the mulloway caught by recreational fishers in the GAB area in 2000 and 2001 were released. Of concern is the possible mortality rate of hooked and released fish. For example, in New South Wales, field and laboratory experiments that were undertaken to determine the mortality rate of mulloway that are hooked and released by anglers, showed that fish with ingested hooks suffer a high mortality rate. Mortality can be reduced by cutting the line rather than trying to remove the ingested hook. During those studies, fish which were mouth-hooked suffered much lower mortality rates, regardless of whether the hook was removed or the line was cut. Cutting the line underwater (and thus reducing the exposure of the fish to air) also reduced mortality rate (Butcher et al., 2007). As indicated in section 2.1.1 Recreational Fishing, mulloway is one of the primary targets for anglers in the GAB, including local fishers, and interstate and overseas visitors. The population of mulloway at the Head of the Bight was also targetted heavily by commercial fishers during the 1990s. At that time, catches and catch rates were reported to be initially high, and then declined significantly as the numbers were exhausted (K. Evans, SARDI, pers. comm., 1996). In the Great Australian Bight (and in other South Australian waters, excluding the Coorong), the minimum legal length for capture of mulloway is 75cm measured from tip of snout to tip of tail, and there is a personal daily bag limit of 2, and a daily boat limit of 6 mulloway (PIRSA Fisheries web site, October, 2009). However, the numbers of mulloway in the GAB have not been determined, and there is a lack of long-term monitoring of total catches from the area. This is concerning, and should be addressed as soon as possible. Western blue groper also occur in the Great Australian Bight. This is a species of conservation concern in South Australia, due to its large size and slow moving nature (and ease of capture), site association, longevity, slow growth, reproductive habits, and population structure (Shepherd and Baker, 2008; Baker, 2009). There is some evidence that, as in some other parts of South Australia, fishers in the GAB cut up juvenile groper (which are commonly called ―green wrasse‖ or ―green parrotfish‖) when they are caught, and use them as bait for catching bigger fishes, and sharks (e.g. Wilson, 2007). Other species of conservation concern, caught on offshore reefs in the GAB area (e.g. off Fowlers Bay, east of the AW NRM boundary) include harlequin fish and southern blue morwong (queen snapper). Harlequin fish populations may be threatened in South Australia because (i) this species, which occurs only in Western Australia and South Australia, has a limited known depth range in upper continental shelf waters, and research, observations and fisheries data indicate that abundance is much lower in South Australia (the edge of the range) compared with Western Australia; (ii) it is a site-associated coastal reef fish with vulnerable population characteristics, similar to other species in the Serranidae family (including solitary nature of the fish, benthic existence and strong site association with reefs and caves, relatively slow growth, large size, and inquisitive nature); (iii) there is a paucity of information about population sizes, biology, and population dynamics (although some aspects of the biology are now being 58 examined by DEH in South Australia: research by S. Bryars); (iv) harlequin fish is caught recreationally and (to a lesser extent) commercially across its range by a number of methods, with few controls on the capture, particularly in South Australia, and (v) insufficient investigation of the potential impacts of fishing has been undertaken to date (summary from Baker, 2009). Similarly, southern blue morwong is a large (to 1m), relatively long-lived species (to at least 21 years) (Coulson et al., 2007), which is vulnerable to over-exploitation and depletion due to its strong site-association with reefs; the migratory and aggregative nature (including spawning aggregations); relatively late age at maturity (3 to 7 years); the existence of inshore nursery areas; and the popularity of this species as a fishing target, for sports fishers / offshore charter boat fishers (summary from Baker, 2009). Large, reef-aggregated morwong are easy to find, and to capture, but there is insufficient information on population sizes over space and time. A number of shark species of conservation concern are caught in State-managed waters of the GAB, including the AW NRM area. Examples include bronze whaler, black whaler (dusky shark), school shark and hammerhead (see Section 2). Although some of the other shark species taken are more abundant over their range (e.g. gummy shark) it is noted that all sharks are potentially vulnerable to population decline due to fishing, because of their low fecundity and reproductive output relative to most bony fishes. Bronze whaler may be vulnerable to over-exploitation due to its behaviour (seasonal aggregations, and use of shallow waters as pupping and nursery areas), longevity (more than 30 years, possibly to 50 years), delayed maturity (17-23 years), biennial reproductive cycle and long gestation (15-21 months) (Walter and Ebert, 1991 and Jones, 2008, cited by Baker et al., 2008). This species is listed by IUCN as Near Threatened at a global scale (IUCN, 2009; Appendix 6). The catch of the related species black whaler / dusky shark C. obscurus is not recorded, and is likely to be low in the GAB; however, the vulnerable status of this species both globally and across southern Australia is noted. During the early-mid 2000s, a report on the status of C. obscurus in Western Australia suggested depletion of the breeding population and declining recruitment, based on apparent reduction in catch rates of neonates (Department of Fisheries, W.A., 2005). Researchers consider that there might be fewer adult females pupping over a smaller geographic range, and this is supported by a recent stock assessment, which concluded that demersal gillnet fisheries‘ catch of primarily neonate (first year) sharks was sustainable as long as mortality of sharks older than 6 years was less than 4%. However, Department of Fisheries research data show that there is a continuing bycatch of adult C. obscurus in other fisheries, and also mortality from entanglement in plastic packing straps. Thus, the collective mortality of dusky sharks is beyond that generated by the managed shark fisheries, and this remains a major cause for concern in W.A. (Department of Fisheries, W.A., 2005; McAuley, 2005), where the fishery for this species is classified as overfished (Larcombe and Begg, 2008). This species is listed by IUCN as Vulnerable at a global scale (IUCN, 2009; Appendix 6). school shark is another shark species caught in the shelf waters of the GAB, as discussed in other parts of this report. However, the catches in the GAB are much lower than in other parts of south-eastern Australia, such as southern Eyre Peninsula, southern and south-eastern Kangaroo Island, south-eastern South Australia and Tasmania (McLoughlin and Wood, in Wilson et al., 2009). During the early 2000s, an assessment of school shark stocks suggested that the biomass of mature school sharks was between 9% and 14% of original levels, and that 59 at current exploitation levels, the species was unlikely to rebuild to its target level by 2011 (Caton, 2003). More recently, available information indicated the pup production is still below 20% of initial levels, and that recent fishery data are a poor indicator of the stock rebuilding progress. The species is still classified as overfished in southern Australia (McLoughlin and Wood, in Wilson et al., 2009), and was listed in 2009 under the EPBC Act 1999 as conservation dependent. There is a report that some relief from fishing pressure on school sharks may occur in coming years as industry reports that fishers are directing their effort away from this species, both because of the need for the stock to recover and because processors are paying less for school shark than previously (Caton, 2003, cited by Trinder, 2007). While this may produce a positive outcome for school shark, it is also likely to lead to increased pressure on gummy shark stocks (Trinder, 2007), although the latter species is currently not considered to be as threatened as school shark in southern Australia (McLoughlin and Wood, in Wilson et al., 2009; IUCN, 2009). Catches of hammerhead sharks in the GAB are not well quantified, but it is noted that the seasonal aggregations of this species, and the use of shallow waters as nursery areas may increase the vulnerability of hammerheads to over-exploitation by fishers. Also, this species is relatively long-lived (i.e. more than 20 years), and matures at a late age (up to 11 years old, for females) (Joung et al., 2005) On a global scale, smooth hammerhead is listed as vulnerable (IUCN, 2009), but it is noted that the fisheries primarily responsible for population depletion are not in South Australia. Nevertheless, the vulnerable status of the species should be considered across the range. Regarding fishing for invertebrates in the GAB, although catches of both abalone and rock lobster are low, and form a small proportion of the region catch of the entire west coast of South Australia, it is noted that stocks of both greenlip abalone and southern rock lobster are in decline in western South Australia. Monitoring is not undertaken in the most remote region of the fishery (i.e. Head of the GAB), but a recent stock assessment for greenlip abalone further east showed that the declines in catch, catch per unit effort (CPUE) and the mean size of abalone in the commercial catch are consistent with declining stock abundance in Region B of the Western Zone (Chick et al., 2007). Blacklip abalone stocks appear to be stable in the area, but decreases in mean size during one recent survey in Region B suggest that closer monitoring is required (Chick et al., 2007). Although the catch of southern rock lobster in the AW NRM is very low compared with other parts of the Northern Zone (see section 2.2), it is noted that a recent assessment (Linnane et al., 2008) reported that (i) catch is decreasing, even with increased fishing effort, in all major regions of the Northern Zone, and the annual quota was not caught in the any of the 5 years to 2007; (iii) catch per unit effort has declined over the past 12 years in all parts of the Northern Zone, and was the lowest on record in 2007; (iv) despite periodic ―good‖ years of juvenile settlement, the adult biomass in 2007 was estimated to be the lowest in the history of the fishery; and (v) the management plan objective of rebuilding the biomass of southern rock lobster was not being achieved. Great Australian Bight Trawl fishing: Although the GAB trawl fishery operates in deeper, Commonwealth-managed waters offshore from the AW NRM region, a number of the fishes caught in the shallowest waters in which the fishery operates, also occur in the AW NRM region. Therefore, the GAB trawl fishery, which is a component of the SESSF (AFMA, 2008c, is discussed here. Over 100 species of bony fishes, sharks, rays and skates are commonly taken 60 in the GAB trawl fishery (AFMA, 2008c), and a similar number of species (presently considered to be of no commercial value) are commonly discarded. The total number of different species observed to be caught during the period 2000-2006 was 282, with many of these species either rarely caught or caught in low numbers. A number of the species taken as by-product are of conservation concern, such as the bronze whalers sharks (see above in section 3.1.1), a species which also occurs in the inshore waters of the AW NRM. Apparently, several tonnes per year are taken by the GAB trawl fishing in the shelf waters of the GAB (Figure 3.11 in Jones, 2008). In a draft Ecological Risk Assessment (ERA) for species in the Great Australian Bight Trawl Fishery (Daley et al., 2006), two of the target fish species taken by the fishery were ranked as ―high risk‖ species, in terms of population impacts from capture in the GAB Trawl Fishery. One example was the Bight redfish Centroberyx gerrardi. Although this species is fished mainly in waters 80m – 210m (Daley et al., 1998) or 150m – 250m (CSIRO et al., 2001) and can occur deeper than 400m, it is discussed here because (i) some operators in the GAB fishery have fished in shallower, mid-shelf waters in recent years (see section 2.2 Commercial Fishing), and (ii) Bight redfish also occurs in shallow waters in South Australia (e.g. 3m - 8m: Australian Museum record; 10m: CSIRO, 2001). A recent review of the status of Bight redfish indicated that this species is of conservation concern in South Australia due to the following points (see chapter on Berycidae in Baker 2010, and references therein):       it is a long-lived (65 - 70+ years) species with high variability in size at age, probably with low natural mortality rate and delayed maturity / high age at first spawning; it is commonly accepted that populations of such species cannot tolerate high and sustained catches over space and time; it is an aggregating / schooling species that is highly vulnerable to capture, and heavily targeted by numerous commercial and recreational fisheries across the range, with a recent age estimate of commercially caught specimens ranging from 9 to more than 60 years; year class strength may be highly variable, and environmentally-driven; despite poor knowledge of the population dynamics, and biomass (for which figures regularly and significantly fluctuate), and inadequate regulation of catches from all fisheries combined, the Commonwealth-managed commercial catches (Great Australian Bight) and commercial catches in W.A. have increased significantly during the past decade, and this species is also an increasingly popular recreational target in S.A. and W.A., particularly charter boat catches; given the size range of commercially caught fish (e.g. in the GAB), compared with the reported size at maturity, it is possible that a significant proportion of the commercial catch is reproductively immature, and has not had opportunity to spawn before being caught; there are indications that the size (and age) structure of the population is being adversely affected by commercial fishing, with a decrease in the modal size of fish observed over 5 years, few larger (older) fish in the catches, and a decrease in the age range of caught fish over nearly two decades (from predominately 20-40 year old fish in the late 1980s, to mainly 14-24 year old fish in the following decade); 61  despite annual abundance surveys using trawls in the Great Australian Bight, there is little information about this species across the range, including reliable measures of relative abundance and biomass over space and time; distribution and connectivity between populations; habitat preferences of adults and juveniles; reproduction; larval movement; population dynamics and ecology. In the GAB trawl fishery, bycatch data gathered by onboard-observer programs show modest discarding of commercial species (summarised in chapters on Berycidae and Platycephalidae in Baker, 2010), but significant discarding of non-commercial species in continental shelf waters, the latter equating to about 44% by weight of the overall catch. Most shelf trawls are reported to be on soft, sandy substrates with little sessile fauna or flora (AFFA, 2004), but some ‗exploratory‘ shots near the established grounds contained significant benthos. Estimates of retained and discarded catches were made through an Integrated Scientific Monitoring Program (ISMP) during the 2000s. The estimated average annual catch (weight) for the period 2000-2006 was 6,881 tonnes. Of this, 4,054 tonnes was retained, and 2,827 tonnes was discarded, which equates to about a 40% discard rate (annual averages) (AFMA, 2008c). Although the bycatch consists of a large number of species, the bulk of the catch by weight comprises small latchet (Appendix 5), sponges, stingarees, and ocean leatherjacket. During 2000 to 2002, studies of bycatch in the central zone of the GAB trawl fishery showed that latchet dominated the discarded catch (47%) with lesser amounts of wide stingaree (17%), ocean leatherjacket (12%) and sponge (9%). In the Central Zone, the mean catches of target species such as deepwater flathead and Bight redfish were similar (approx. 320 kg per shot), whilst the discarded Latchet was about twice this quantity (approx. 660 kg per shot). Only 0.3% (individuals > 35cm caudal fin length) of the estimated 21 tonnes of latchet landed was retained, i.e. 99.7% of the latchet catch was discarded because it was not considered of commercial size. During the mid-2000s, ISMP data indicated that, on average, 6 tonnes per trip of latchet were discarded, but this was less than in previous years, because fish larger than about 30cm are now retained as by-product. During the 2000-2002, monitoring of bycatch in the central zone of the GAB trawl fishery, about 75% of the landed catch of ocean leatherjacket was discarded, with a mean discarded catch of 165kg per shot compared with a mean retained catch of 54kg per shot. Only large ocean leatherjacket (>33cm Total Length) were retained (Brown and Knuckey, 2002). Smaller ocean leatherjackets (<30 cm TL) have been routinely discarded in the past. Integrated Scientific Monitoring Program (ISMP) data indicate that extremely large catches of juvenile leatherjackets (3-4 tonne per shot) can be avoided by not trawling in shallow waters (100-120m) during the full moon. Other species are discarded in significant quantities, with examples including wide stingaree (Urolophus expansus), spikey dogfish / shortnose spurdog (Squalus megalops), rubyfish (Plagiogeneion macrolepis), ringed toadfish (Omegophora armilla), deepwater burrfish Allomycterus pilatus, various rays and skates, and catsharks (see Brown and Knuckey, 2002). These species occur in waters deeper than the AW NRM boundary, and therefore are not discussed here. It is noted that in 2006, during a draft ecological risk assessment for species in the GAB trawl fishery, various other species either taken as by-product, or discarded in the bycatch - including jackass morwong, yellow-eye snapper, swallowtail, mirror dory, king dory, blue-eye trevalla, spotted warehou, Australian tusk, bigscale rubyfish, pink ling, gemfish and latchet - were all listed as being at high risk of population impacts from the operation of the Great Australian Bight Trawl Fishery (Daley et al., 2006). Gemfish and blue-eye trevalla are 62 classified as species of low resilience, with gemfish being highly to very highly vulnerable and blue-eye trevalla being moderately to highly vulnerable to fishing-induced population decline (Cheung et al., 2005; Froese and Pauly, 2009). In the draft Ecological Risk Assessment (ERA) for species in the Great Australian Bight Trawl Fishery (Daley et al., 2006), 39 by-product species and 15 by-catch species were ranked as being at ―high risk‖. However, in a residual Risk Assessment of the Level 2 Ecological Risk Assessment, the two target species at high risk were considered at lower risk than when initially assessed, due to management measures, such as the introduction of a catch quota for Redfish in 2006 (despite annual increases in quota since that time). As part of that Residual Risk Assessment, the numbers of by-product and bycatch species considered at ―high risk‖ were reduced to 27 and 12 respectively, compared with the initial ERA. However, in a ―rapid quantitative Level 3 assessment, or Sustainability Assessment of Fishing Effects (SAFE) assessment‖ undertaken by Australian Fisheries Management Authority, a supplement to the ecological risk assessment (by CSIRO) for the GABTF, all target, bycatch and by-product species were ranked as ―low risk‖ of impact from operation of the fishery. According to AFMA (2008b), the Level 3 assessment process considers the mitigating effects of management arrangements that were not explicitly included in the ERAs, or introduced after the process commenced. It is noted that the level 3 assessment is made proportional to the spatial area in which the fishery operates, compared with the spatial area of distribution of the species, but this does not account for aggregation, rather than even distribution throughout space (the latter of which is unrealistic, e.g. for reef-associated species in a heterogeneous benthic environment). It is unlikely that the Level 3 assessment provides an accurate indication of fishing-induced threats to marine species in the GAB. There are few reported interactions of the GAB trawl sector with seabirds. Although ISMP data indicate that albatrosses and shearwaters follow the fishing vessels in the GAB (e.g. over 300+ shearwaters and <10 albatross followed one of the 10 GAB trawl boats when it was monitored in spring and summer of 2006), no mortalities were observed in 132 trawl shots of that period, only one seabird mortality was observed (a fleshy-footed shearwater drowned after being entangled in the trawl net). More albatrosses are observed in early winter, with the start of the annual northern migration of albatross. Mortality of syngnathids (seahorse, seadragon, pipehorse and pipefish family) is also reported to be low in the GAB trawl fishery. During monitoring of one vessel in 2006, one spiny pipehorse (Solegnathus spinosissimus) was reportedly found on the deck after trawling over ―weedy‖ bottom (ISMP data, 2006), but it is noted that this south-eastern Australian and New Zealand species does not normally occur in South Australia, and the specimen might have been misidentified. Note that all pipehorse species are protected under legislation in South Australian waters under the Fisheries Management Act 2007, and are considered uncommon in South Australia, and potentially at risk (Baker, 2009). A draft ecological risk assessment of the GAB trawl fishery during the early 2000s (Daley et al., 2006) identified 6 high risk habitats on the outer shelf (100m – 200m) trawled by the fishery. These were mainly soft sediment seabed types characteristically dominated by large sponges and mixed epifauna, with bryozoan communities at the shelf break. Sedimentary, sub-cropping rock with communities of large sponges also scored at high risk. Slope habitats, some also classified as high risk of trawl damage, are not discussed here, because they are much deeper than the boundary of the AW NRM region. 63 The bycatch of sponge in part of the GAB trawl fishery has been quantified in some years. For example, during a bycatch survey from 2000 to 2002, Brown and Knuckey (2002) recorded sponges in 131 or 209 trawl shots (i.e. more than half), with an average discard rate of 9.5kg of sponge per shot. During the monitoring period, sponge constituted 9% of the total discarded bycatch from the central zone (deeper Commonwealth-managed waters seaward of the AW NRM boundary). Hard coral was recorded in 24 of 209 shots (average discarding of 50kg per shot). The bycatch of sponge, rock and hard coral raises concerns over the adverse effects of trawling on benthic communities in the Great Australian Bight (Brown and Knuckey, 2002). Attached benthos is a major component of benthic habitat ‗structure‘ which is important for maintaining community and ecosystem functioning, because of the associated quantity and variety of food resources, living spaces and refugia. Habitat structure provides refuge from competitors, unfavourable environmental conditions and predators, which is particularly important for the early life stages of fish (Brown and Knuckey, 2002). Factors that affect juvenile survivorship influence the size, vitality and distribution of adult populations, so habitat complexity has an important role in the sustainability of fish stocks (Brown and Knuckey, 2002). Sponge beds can take many years to recover from damage. Other Fishing in the GAB: In deeper shelf waters in the south-eastern GAB (i.e. not in the AW NRM region), southern bluefin tuna (SBT) Thunnus maccoyii is caught in a Commonwealthmanaged fishery. This species occurs in both shelf and slope waters along the Western Australian coast and the Great Australian Bight (CSIRO data, 1979; Kailola et al., 1993; Gunn and Young, 1999; Wilson et al., 2009). The small number of tuna taken by charter boats and other recreational fishers in shallow water is not significant, compared with the number (99% of the Australian catch) taken by commercial fishers in the south-eastern GAB. The Head of the Great Australian Bight may be significant for juvenile southern bluefin tuna (Figure 22 in Ward et al., 2008). Aerial surveys along the shelf and slope of the central and eastern Great Australian Bight, showed that over the period 1993 to 2007, the number of sightings declined, from approximately 130-180 sightings per annum (reportedly representing about 12,000 – 22,000 tonnes of SBT per annum) during the early-mid 1990s, to around 40 – 80 sightings (approximately 4,000 – 6,000 t) during the mid-late 2000s (Eveson et al., 2007). Southern bluefin tuna is an overfished species, of high conservation concern, and is currently listed under the IUCN Red List as Critically Endangered (IUCN, 2009). In Australia, it is classified as overfished (Wilson et al., 2009), but not formally protected under the EPBC Act 1999 as a threatened species, because the nomination for listing was not accepted by the Australian government (Department of the Environment and Water Resources, 2005). Other issues associated with the capture of southern bluefin tuna in south-eastern GAB include anecdotal reports of deliberate harm to sharks and marine mammals which interact with the tuna fishery, e.g. which follow the captured tuna that are being towed eastwards to pens in Spencer Gulf. Kemper et al. (2005) provided a brief summary for locations east of the GAB (e.g. Port Lincoln area), where deliberate harm of cetaceans has occurred. Fishing Threats to Sea Lions: In the deeper shelf and slope waters away from the AW NRM region, bycatch and incidental injury and mortality in fishing operations may be an ongoing and significant issue of the Australian sea lion, which feed in the area (Goldsworthy et al., 2009a; Hamer et al., 2009). In recent decades, low levels of bycatch of Australian sea lions have been reported by a number of commercial fisheries, including the western rock lobster fishery (WRLF) in W.A., but captures are highly variable both within and among fishing seasons (Gales and Wyre unpubl. report., Mawson and Coughran, 1999; Shaughnessy et al. 2003, cited by 64 Campbell, 2005). Video evidence of Australian sea lions around rock lobster pots in Western Australia has shown that individual animals can consume up to five lobsters from a commercial pot within 2 hours, and a group of 8-10 animals consumed 23 lobsters from a single pot in approximately 3 hours (Campbell, unpubl. data, cited by Goldsworthy et al., 2007). This rate of consumption and interaction suggests there may be a considerable trophic interaction of pinnipeds with rock lobster fisheries. Additionally, there is a high incidence in some locations of young sea lions becoming entrapped in lobster pots (Goldsworthy et al., 2007). Higher levels of bycatch have been recorded in the shark fishing component of the Gillnet, Hook and Trap sector of the SESSF in South Australia, and this has long been an issue. In South Australia, the entanglement of sea lions in monofilament netting of 150 mm mesh (used in the shark fishery) has been described at least as far back as the 1980s (e.g. Robinson and Dennis, 1988). Anecdotal reports from shark fishers suggest that entanglement of sea lions occurs in inshore rather than offshore waters (Shaughnessy et al., 2003, cited by National Seal Strategy Group and Stewardson, 2005). More recently, a study on the bycatch of Australian sea lion in shark gillnet fisheries operating in South Australian State- and Commonwealth-managed waters has confirmed that this may be a significant source of mortality for sea lions in South Australia (research by Goldsworthy et al., SARDI Aquatic Sciences; Goldsworthy and Page, 2007; Goldsworthy, 2008). Readers are also referred to the report by Hamer et al. (2009) for further information, which was not available for citation at the time of writing (S. Goldsworthy, SARDI, pers. comm., 2010). An additional if small source of mortality may be shooting of sea lions, reportedly from onboard fishing vessels (e.g. in W.A. – Campbell, 2005). Although it is unlikely that the removal of a small number of animals in an isolated instance would have a significant impact on population size or viability, the combination of direct shooting and harassment could disrupt a small breeding colony (Campbell, 2005), but data are lacking from remote areas such as the Great Australian Bight. The high level of subdivision within the Australian sea lion population, coupled with strong natal site fidelity of females, means that recruitment of females is solely from within each individual colony. The combination of increased mortality and the potential effects of reduced genetic diversity among small separated sub-populations may limit their recovery rate and/or viability (Amos and Balmford, 2001, cited by Campbell, 2005). Fishing Threats to Cetaceans: The interaction of whales and dolphins with fisheries that operate in deeper GAB waters seaward of the AW NRM has not been fully quantified or publicised. It is noted that cetaceans are not included in bycatch surveys (e.g. Brown and Knuckey, 2002) or risk assessments for the GAB trawl fishery (e.g. Daley et al., 2006; AFMA, 2008a,b,c). However, it is noted that dolphin injuries and mortalities have occurred in the shark fishery component of the SESSF, operating in other parts of South Australia (data by C. Kemper, South Australian Museum). Entanglements can be fatal, especially when small cetaceans are caught in nets, or non-fatal. Non-fatal entanglements can result in reduced reproductive potential, and can interrupt migration, and increase the risk of re-entanglement (Kemper, 2004). Competition between fisheries and cetaceans for prey is another issue, in areas where significant quantities of scale fishes and squid are taken by trawl and hook fisheries, but this issue is not relevant in the shallow waters of the AW NRM region. Common Dolphins consume bony fishes more than cephalopods (squid and cuttlefish), and bottlenose dolphins consume more cephalopods (octopus, cuttlefish, squid) than bony fishes (Gibbs, 2004). The deliberate harming of cetaceans by fishers, which has been documented in other parts of South Australia (e.g. Kemper et al., 2005), has not been investigated in the GAB. 65 Threats to Sea Lions The threat that fishing operations in deeper waters (seaward of the AW NRM) may pose to Australian Sea Lion populations is discussed in the previous section. Historically, populations of the endemic Australian Sea Lion were greater in both W.A. and South Australia; however it is likely that sub-populations in the Great Australian Bight were never commercially harvested due to the isolation and general inaccessibility of the coast from both from land and sea (Dennis and Shaughnessy, 1996). It is notable, however, that analysis of historical patterns of abundance on the west coast of Western Australia suggests that the population size was greater prior to the impacts of colonisation and commercial sealing/whaling between the 18th and 20th centuries (Campbell, 2005). The major historical impacts on Australian Sea Lion populations as a whole were a combination of subsistence and commercial harvesting events from the 1700s to the 1920s (Campbell, 2005). The GAB represents a significant barrier to dispersal (Campbell, 2003, cited by Goldsworthy et al., 2009a) and the Bunda Cliffs population may therefore be effectively isolated. It is likely that sea lions from these colonies were not previously harvested to the extent that colonies in other parts of S.A. and W.A. were harvested. The current status of the breeding population in the GAB is unclear (Goldsworthy et al., 2009), due to lack of regular monitoring. Threats to Australian Sea Lions within the AW NRM are relatively low compared with some other parts of South Australia, due to the legislative protection afforded by the Great Australian Bight Marine Park, and by the Nullarbor National Park (which abuts the coast and ostensibly protects colonies on coastal cliffs, rocks and caves from harm). There is a low level of human disturbance to breeding colonies and haul out sites, due to inaccessibility. However, it is noted that Edyvane and Andrews (1998) reported that two of the Bunda Cliff sites, located on the Yalata Aboriginal Land Lease near Twin Rocks, in addition to sea lion breeding colonies at highway viewing points along the Eyre Highway (located 40km and 125km from the Western Australia border) are situated directly below the car park areas. According to Edyvane and Andrews (1998), these colonies may have been subjected to disturbance from above, as evidenced by the quantity of rubbish recorded in the colonies. Such items included car and truck tyres, highway signs and bottles. The authors considered that such materials may cause direct injury to the animals or at least some disturbance. In deeper waters, impacts that may affect Australian sea lions include:   fisheries entanglements (e.g. shark gillnets) – see previous section, and competition with fisheries for common prey species (sea lions are known to rob baits from lobster pots, as well as fish in nets set for sharks). Given the time that sea lions spend searching for food in deeper waters, interactions with . fishers may occur where foraging area and fishing area overlap. Recommendations from fishery Ecologically Sustainable Development (ESD) assessments, fishery Bycatch Action Plans (BAPs), and a recently drafted Commonwealth government Draft Recovery Plan for the Australian Sea Lion, have all identified the importance of assessing and mitigating interactions between sea lions (and fur seals) and commercial fisheries. An example of such fisheries interaction in South Australian waters is the attraction of sea lions (and seals) to fish trapped in set gill-nets, and consequent entrapment, resulting in the animals either drowning, tearing out a section of net, or being cut free by fishers (Robinson and Dennis, 1988; Shaughnessy et al., 66 2003, cited by National Seal Strategy Group and Stewardson, 2005). Although there are few official records of seal interactions with the shark fishery, there are various anecdotal reports of shark fishers finding drowned seals in demersal shark nets set near breeding colonies, in State waters (Shaughnessy and Dennis, 2002). In 1996, a fisher from South Australia reported to P. Shaughnessy that he caught 20 Sea Lions a year in his shark nets set off Neptune Islands and Kangaroo Island (Shaughnessy et al., 2003, cited by National Seal Strategy Group and Stewardson, 2005). Records of Sea Lions and fur seals entangled in marine debris also indicate that a significant level of interaction occurs with set nets and lost monofilament netting throughout South Australia. In the GAB, Australian sea lions have been recorded entangled in sections of commercial shark net at a number of sites, including Jones Island (Shaughnessy and Dennis, 2002) and Nicolas Baudin Island (J. McKenzie, pers. comm., cited by National Seal Strategy Group and Stewardson, 2005). There is limited information on the trophic interactions between Australian sea lions and commercial fisheries in the Great Australian Bight. Goldsworthy et al. (2003) estimated the spatial distribution of foraging and consumption efforts of seals in southern Australia, and identified regions in the eastern Great Australian Bight around Kangaroo Island and off southern Eyre Peninsula as being areas where consumption was high, and where trophic interactions with fisheries may be significant. Fine scale data on the spatial distribution of commercial fishery catch in the GAB, coupled with information on the foraging area of sea lions when away from colonies, would enable the extent of overlap in fishery catch and seal consumption to be estimated. This could be used in conjunction with dietary and food-web studies, to determine the degree of trophic interactions (Goldsworthy et al., 2007). Other threats to sea lions which may be an issue in some parts of South Australia (e.g. shooting / harassment, oil spill impacts, and disease) are much less likely to affect populations in the GAB, due to the remoteness of the area, and its position away from population centres and industry. Population viability analysis (PVA) on Australian sea lion subpopulations has indicated that large numbers of subpopulations with low pup production are vulnerable to extinction. PVA simulations suggested that in the absence of human-induced mortality, a number of Sea Lion subpopulations will go ―quasi-extinct‖ (<10 females), but in the face of small (1-2 additional females/year) but sustained fishery bycatch-induced mortality, most other small subpopulations will become quasi-extinct and negative growth will become a feature of even the largest subpopulations (SARDI data; Goldsworthy, 2008). In South Australia, there is apparent depletion (i.e. very low pup production) of a large number of subpopulations that may be indicative of widespread subpopulation declines in the species. Recreation and Tourism Compared to other Australian marine areas, there is a relatively low level of human use in the Great Australian Bight (Director of National Parks, 2005). Management issues in the coastal zone that arise from recreational and tourism activities are discussed the Far West Coastal Action Plan and Conservation Priority Study (Caton et al., 67 2007), and thus will not be reiterated here, other than to note that such issues include wildlife disturbance, vegetation destruction, dune de-stabilisation, accelerated cliff erosion in some places, soil disturbance and compaction, weed introduction and litter. Many of these impacts are concentrated around formal and informal car parks and camping areas (Caton et al., 2007). Off-road vehicle tracks have been identified as a significant threat throughout the area, and mapping undertaken by Segaran (2006, cited by Caton et al., 2007) showed a large increase in the number of tracks from 1979 to 2004. Open access to the cliff areas may be reducing the nesting success of osprey and white-bellied sea eagles (Caton et al., 2007; Dennis, 2008). Both species are considered extremely vulnerable to human disturbance. Dennis (2008) reported that an abandoned osprey nest on a pinnacle below cliff-line at 697353E 6513730N (~4km west of the Head of Bight visitor centre), may have come under the same disturbance pressure from increasing human activities in the Head of the Bight area as a former white-bellied sea eagle site, almost two decades earlier. In 1994, an abandoned white-bellied sea eagle nest structure was recorded ~2km east of the Whale Sanctuary Visitor Centre and viewing platforms. At that time the annual winter presence of Southern Right Whales was established, and increasing numbers of visitors and whale research team activities in the area are considered to have displaced the sea eagles from the area (Dennis, 2008). As for the abandoned osprey nest cited above, there is a frequently used vehicle track close to the cliff-edge, directly above the abandoned osprey nest site in the eastern Bunda Cliffs (Dennis, 2008). One of the potentially threatened bird species in the AW NRM region is the hooded plover. Generally, the nesting period extends between August and February in South Australia. It is during this time that the adults and chicks are most vulnerable to disturbance and destruction. The eggs are laid in shallow sand scrapes or seagrass wrack and, if they hatch successfully, the chicks may then expend large amounts of energy, sometimes with fatal consequences, running to hiding places to avoid predation or harm. The young do not fly for at least three weeks, so each clutch is vulnerable for nearly two months. The hooded plover's habit of leaving a nest site if people approach, and usually not returning until people have left the area, has an important influence on breeding success. Caton et al. (2007) discussed locations within and adjacent to AW NRM area in which hooded plovers nest, or may nest. Impacts on hooded plover populations may come from cars driving on beaches (and therefore crushing nests and eggs); humans trampling nests; and human presence near nesting areas, which may disturb adults and/or chicks. The destruction of nests by off-road vehicles has been a major factor in other parts of South Australia (such as the Coorong region). hooded plovers nesting in exposed areas of beaches may also suffer high levels of egg and chick predation by the European red fox, feral cats, and feral dogs (Caton et al., 2007) and domestic dogs, particularly when the parent birds are absent. The increasing number of tracks in the AW NRM coastal area during the past decade, may also lead (or may also have led) to increased visitor access to beaches, including some beaches where hooded plovers nest. Silver Gulls may also prey upon hooded plover eggs and chicks, but this has not been documented specifically for hooded plover nests in the AW NRM region, and it is not known for this report whether such predation occurs. Other species within AW NRM that may be sensitive to disturbance by humans include the Sooty Oystercatcher, Banded Stilt, Common Sandpiper, Sanderling, Eastern Reef Egret, Pied Oyster Catcher, Grey-tailed Tattler and Kelp Gull (Caton et al., 2007). 68 Marine Debris Much of the information on marine debris in southern Australia does not apply to the AW NRM region, and there are few data for the Great Australian Bight (C&R Consulting, 2009). The central GAB is one of the 4 regions in Australia for which there are few records of wildlife impacted by plastic debris (C& R Consulting, 2009). It has been noted that Southern Right Whales are considered at high risk of entanglement due to their tendency to aggregate inshore during travel and calving (Jones, 1995; Allen and Bejder, 2003, cited by C & R Consulting, 2009), but during a recent review for the Commonwealth government, too few records were available for any seasonal patterns to be discerned (C& R Consulting, 2009). A study of the distribution of known records of cetaceans and seabirds impacted by plastic debris (entanglement and ingestion) in Australian waters since 1998, showed that there were 1 to 3 records of cetacean entanglements from the Great Australian Bight (from the Ceduna / Streaky Bay area, east of the AW NRM region), and no records of seabird entanglements from the area (C & R Consulting, 2009). The results likely reflect the lack of monitoring of entanglements in the remote GAB region, particularly the Head of the Bight, rather than genuinely low levels of entanglement. In Anxious Bay in the eastern Great Australian Bight, SARDI Aquatic Sciences ran a long-term litter clearing program on the beach in that area (Edyvane et al., 2004), with the aid of community groups. Although Anxious Bay is not in the AW NRM region, the results of the survey may be relevant, because the prevailing winds and currents are likely to move marine litter eastwards to Anxious Bay, from waters further west in the GAB, where it may have originated from fishing vessels and other marine craft. The SARDI surveys showed that over the 1991–1999 period, a large but gradual decline in the amount of beach-washed litter was recorded (with minor peaks recorded during 1992 and 1994). In a decadal comparison, beachwashed litter decreased from 344kg recorded in 1991 (13.2 kg/km) to 49 kg in 1999 (1.9 kg/km), but reached a maximum of 390 kg in 1992 (or 15 kg per kilometre of beach). A sharp increase in litter was recorded in 2000 (i.e. 252 kg or 9.7 kg/km), reported to be probably due to stronger than average onshore surface flow (or Ekman Transport) in the western Eyre Peninsula and Bight region at that time (Edyvane et al., 2004). Prior to the survey in 2000, the results appeared to indicate that ocean litter on Anxious Bay beach was beginning to level out at around 50–70 kg/year (i.e. 2–3 kg/km). The yields and type of litter collected from the annual surveys indicated that the majority of litter washed ashore originated from commercial fishing activities within the GAB. Most of the fishing-related litter was directly sourced to the Southern Rock Lobster Fishery (i.e. bait buckets, baskets, pots), the Great Australian Bight trawl component of the SESSF (i.e. cod-ends, trawl nets) and the gillnet and shark hook sector of the SESSF (i.e. mono-filament gill-nets and long-lines) (Edyvane et al., 2004). Between 1994 and 1999, large reductions were observed in the amount of bait straps (77% reduction), lobster bait baskets/buckets (86% reduction), nets/ropes (62% reduction) and floats/buoys (83% reduction). According to the results of the SARDI-managed surveys, fishingrelated litter in the Bight has reduced at a slower rate than domestic litter. The level of glass and soft plastics on the beach reduced from 103kg to 7kg and 119kg to 8kg respectively, and the level of hard plastics reduced from 122kg to 30kg. To date, the South Australian Rock Lobster Fishery, the south-east trawl sector, and the gillnet and shark hook sectors of the SESSF are three fisheries which have been highlighted as responsible for significant numbers of interactions with fur seals and sea lions in South Australia 69 (Page et al., 2004; Goldsworthy et al., 2007, 2009, and see 3.1.1 and 3.1.2 above). Previously, bait box straps were one of the most significant sources of marine litter that had detrimental effects on marine fauna. For example, bait straps were the most common material (30%) observed entangling New Zealand fur seals during a 6-year study on Kangaroo Island, and accounted for 11% of material identified entangling Australian sea lions during a 15-year study (Page et al., 2004). Other material included trawl-netting (28%), rope (23%), plastic bags (7%), hooks and fishing line (3%), monofilament netting (1%) and other material including rubber ‗orings‘, string and lobster-pot (8%) (Page et al., 2004, cited by Goldsworthy et al., 2007). Hamer et al. (2009) Goldsworthy et al (2009) discuss fishery interactions in more detail, and outline ways in which interactions can be minimised. Some fisheries (i.e. southern rock lobster, and the gillnet and shark hook sector of the SESSF) have shown marked reductions in fishing-related litter. This may be due to both reduction in fishing effort over the past decade, and also efforts that have been made within marine industries to reduce the amount of marine litter entering the environment in South Australia. For example, the South Australian Rock Lobster Fishery has moved away from potentially harmful packaging materials such as hard plastic bait box bands, in favour of glued boxes (DEWHA, 2009a). At a national level, a threat abatement plan for the impacts of marine debris on vertebrate marine life is being developed (DEWHA, 2009a). Although South Australia is only briefly considered in the plan, many of the proposed actions are relevant to marine debris reduction in all States, and in Commonwealth waters. Adherence to the plan will also ensure that operators in Commonwealth and southern Australian State waters remain compliant with the international MARPOL (Annex V) regulations. Marine Pests According to the National System for the Prevention and Management of Marine Pest Incursions (NIMPIS, 2009, 2010), in conjunction with a national ranking of marine pests in Australia (Hayes et al., 2005) and several other sources combined, at least 54 marine and/or estuarine pests occur in South Australian waters (Table 5 below). The majority of these pests were introduced into South Australian ports and harbours by ships‘ ballast water or as fouling on ships‘ hulls, and it is unlikely that most of these pests have colonised the relatively pristine and wave-exposed waters at the Head of the Bight, where there are no ports or harbours, and relatively undisturbed substrates. The nearest port is at Ceduna, approximately 200km east of the eastern marine boundary AW NRM area. However, due to the remoteness of the AW NRM region, there has been no full investigation of introduced marine species in the area. As a result, introduction of marine species may go unnoticed, and of those that possibly occur, none have been confirmed, and the sites of origin are not known (i.e. these would be classified as ―cryptogenic species‖). Although an interactive mapping function in NIMPIS (2010) reports no marine pests as occurring in the Head of Bight area, the following potential pests are noted here: 70 Hydroids The hydroid Antennella secundaria has been recorded in open waters with good current flow in South Australia (NIMPIS, 2002). It grows on variety of substrates, (including red algae, seagrasses, ascidians, sponges, bryozoans, shells, stones, crustaceans), but there have been no recorded impacts of this species in southern Australia. Bryozoans The globally widespread encrusting bryozoan / sea lace Membranipora membranacea occurs in South Australia (Bock, 1982), but its full distribution here is uncertain. This species attaches to kelp and other organisms, and studies in other countries have shown that it can cause seasonal dieback of kelps, which can aid the establishment and growth of nuisance macroalgae (see Table Y). Its presence in the AW NRM is unknown. The bryozoan Cryptosula pallasiana may have potential to occur in the area, but is mainly found on seagrasses, drift algae, oyster reefs and human-made structures. Colonies found to date in Australia generally do not reach a large size or cover large areas of substrate. The lace coral Schizoporella unicornis can exist on a broad range of substrates. It has been reported in the eastern Great Australian Bight, east of the AW NRM region, and is also known from south-western Australia. The impacts of this species are not known (NIMPIS, 2010). Common fouling organisms on vessels, such as Watersipora arcuata, are now widespread along the southern Australian coastline, and easily transferred. This species is resistant to anti-fouling paints; can spread rapidly on vessel hulls and provide an area for other species to settle upon. Also attaches to rocks, macroalgae and wooden structures. This species has been recorded in the eastern Great Australian Bight, east of the AW NRM region, but is not known to date any further west. Molluscs The New Zealand Creeper Shell or mud snail Zeacumantus subcarinatus is one of the introduced marine species that might occur in the region. It was introduced to Australia from New Zealand during the 1930s, and is now broadly distributed across the southern coast of Australia. It occurs mainly in the intertidal, such as rock pools at high tide level in bays and estuaries, and on the open coast. There are records from Streaky Bay, east of the AW NRM region. This species is known to graze green macroalgae such as Ulva (e.g. McClatchie, 1979). Overseas, much research has been undertaken on trematode parasites that infest this creeper shell. The impacts of introduced populations are not well known (and may be negligible in regions such as the AW NRM, even if present). It is also possible that the Mediterranean Blue Mussel Mytilus galloprovincialis occurs in the AW NRM region, given its presence in bays of the eastern Great Australian Bight, and its culture in South Australia. It is possible that this species came to Australia on hulls of early sailing ships when the continent was colonised by Europeans, and has long been known as the ―native species‖ Mytilus edulis, M. planulatus, or M. edulis planulatus. The impacts of blue mussels in relatively undisturbed areas such as the AW NRM region are unknown, but if this species is present, its impacts may not be serious, compared with harbour areas and disturbed substrates where Mytilus can compete with other organisms for space and food. A beach clam from New Zealand, Paphies (Mesodesma) ventricosa was reported by the Academy of Natural Sciences (2006) to occur in the Great Australian Bight, but there is very little information, and it is not listed in documentation for marine pests in southern Australia (e.g. Hayes et al. 2005; NIMPIS 2010). 71 According to Department for Environment and Heritage SA (2001), the New Zealand Screw Shell Maoricolpus roseus has been recorded from the ―west coast of South Australia‖, but NIMPIS (2010) and other references report it only from south-eastern Australia, where it was introduced to Tasmania in the early 20th century, and is now firmly established in Tasmania, Victoria and New South Wales. It has the potential to spread westwards. The impacts of dense populations of M. roseus on native species have yet to be determined, but could possibly reduce numbers of native gastropods and bivalves via competition for food and space. There appears to be no evidence to indicate that this species occurs yet in the Great Australian Bight. Hayes et al. (2005) reported that the shipworm Teredo navalis might occur in the Eucla marine bioregion of South Australia (of which AW NRM area is part), based upon the generalised distribution of this species across southern Australia, and its temperature tolerance. This species bores into wooden structures such as boats and harbour piles, but is unlikely to become a pest in the AW NRM region, even if present. Feral populations of the Pacific Oyster Crassostrea gigas are found in the Ceduna area (NIMPIS, 2010), due to the prevalence of oyster culture in the bays of eastern Great Australian Bight. Presence in AW NRM region is unknown. Worms The Slime Featherduster Worm Myxicola infundibulum might occur in the eastern Great Australian Bight, but there are no confirmed records. It is transported by vessels, and there are no recorded impacts. Crustaceans The Acorn Barnacle Megabalanus tintinnabulum, is a common fouling species that is widespread in Australia and found in many types of habitat. There are no records from the Great Australian Bight, but it has potential to spread to the area. Ascidians The colonial ascidians Botryllus schlosseri and Botrylloides leachi may both occur in the AW NRM area, but there are no confirmed records. One has been recorded as far west as the Murat Bioregion, and the other as far west as the Eyre Bioregion (see Table 5). Both can attach to rocks, and to a variety of benthic marine organisms, as well as vessels, and human-made structures associated with harbours and aquaculture. Colonies are eaten by crabs, gastropods, nudibranchs and flatworms. Can be transferred by vessels, fisheries activity and aquaculture. Although colonial ascidians can be dominant competitors, overgrowing and excluding many other fouling species, the impacts are low compared with many other marine pests. Several toxic dinoflagellates are of concern in South Australia (see Table 7), but are mainly associated with ports, and estuarine areas, and unlikely to cause significant harm in the AW NRM region. According to a recent analysis of South Australian Herbarium data (Baker and Gurgel, in prep.), one cryptogenic species of macroalgae occurs in the AW NRM region, the green alga Ulva rigida. This species is cosmopolitan in distribution; widespread Europe and Africa (> 15 countries each); also Atlantic islands; Nth, Central & South America; SW Asia; Philippines; Australia, New Zealand & Antarctica (Cowan, 2006; Guiry & Guiry, 2010). In Australia, it is known from most States (SA Herbarium data; Baker & Gurgel, in prep.). Many Ulva species, including U. rigida have the capacity to bloom in abundance, particularly in eutrophic conditions, 72 such as in estuaries, port and marinas (e.g. Sfriso, 2010); however, the chance of this species becoming a nuisance in the AW NRM region is low. Of significant concern in Gulf St Vincent is the presence and encroachment (e.g. into seagrass beds) of the invasive green macroalga Caulerpa taxifolia, and concerted research and management efforts have been made to control this species during the past decade (e.g. Cheshire et al., 2002; Rowling, 2007; Deveney et al., 2008; Wiltshire and Rowling, 2009). However, it is noted that Caulerpa taxifolia has a relatively low probability of invading remote, low nutrient and wave-exposed areas such as AW NRM, unless boats or associated equipment (e.g. anchors) from upper Gulf St Vincent with fragments of Caulpera on them are permitted to enter AW NRM waters. This species can proliferate in warmer waters (22–25°C) compared with cool waters (15–18°C) (Glasby and Gibson, 2007), and waters at the Head of the Bight are seasonally warm; therefore caution is required. Two of the most serious marine introductions in southern Australia have not yet reached South Australian waters. These are (i) the Japanese seaweed Undaria pinnatifida, which can grow to 2m in height, and was introduced to Tasmania by shipping, and is now also present in Victoria. Undaria grows rapidly and can shade the canopy on reefs, killing other macroalgae, and this can adversely affect the ecology of the reef system, particularly for native herbivorous animals; and (ii) the northern Pacific seastar Asterias amurensis, a highly fecund species which reproduces prolifically, is a voracious predator, and is a particular threat to native shallow-water communities of mussels, scallops, oysters and clams, and to shellfish aquaculture facilities. It also preys on other seastars, and will scavenge dead fish. Asterias was introduced to Tasmania by Japanese shipping, and has spread to Victoria (NIMPIS, 2002). It is found mainly in the shallow subtidal, but can also live in deeper shelf waters. It has few known predators (none of which are in Australia), and larvae are easily transported in ships‘ ballast water. The Centre for Research on Introduced Marine Pests (CRIMP) at CSIRO in Hobart, is conducting research into using a biological agent as a long-term control method. Although it has not yet been recorded in South Australia, it is considered likely to spread from south-eastern to south-western Australia if populations are not controlled. Fortunately, this species does not normally occur in exposed, high wave energy areas, and thus the likelihood of a population becoming established in the AW NRM region is low. There is a possibility, given the likely impacts of climate change on ocean warming and ocean circulation patterns, of several marine pest species in South Australia becoming more widespread in future (see 3.1.8: Climate Change). 73 Table 7: Marine pests reported in South Australia (according to Bock, 1982; McMinn et al., 2000; Department for Environment and Heritage, 2001; Russell-French, 2002; Hayes et al., 2005; Department of Fisheries W.A., 2005; Reynolds, 2005; Academy of Natural Sciences, 2006; PIRSA, 2007; Department of Health, 2009; NIMPIS, 2002, 2009, 2010; Wiltshire and Rowling, 2009; ABRS, 2009). Other references include Dixon et al., 1981; Gunasekera et al. 2005; Department of Fisheries WA, 2005a, 2005b; Beechey, 2009; Scheibling and Gagnon, 2009). GSV = Gulf St Vincent. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Macroalgae Caulerpa racemosa var. cylindracea Caulerpa Unlikely Reported in South Australia from GSV bioregion (e.g. St Kilda, North Haven and O‘Sullivans Beach boat ramps). Caulerpa taxifolia aquarium Caulerpa Unlikely (but caution required) In South Australia, mainly known from Port River, West Lakes and associated areas in north-eastern GSV. Highly invasive, and found on a wide variety of substrates, including rock, sand and mud to dead seagrasses, mainly in waters less than 15m (in Australia). The invasive aquarium strain can occupy up to 100% of the available substrate. Native Australian populations in tropical waters are found on rocky reefs and seagrass meadows in sheltered or moderately wave-exposed areas in both polluted and pristine waters. Easily transferred on vessels hulls and fishing equipment etc. Polysiphonia brodiei a red macroalga Unlikely To date, known in South Australia from GSV, and the upper south-east (e.g. Robe area). In Australia, found to date mostly in port environments. Colonises wooden structures such as jetties and pylons, floating structures such as ropes, buoys and vessels and other fouling species such as mussels. However, note that the species is easily transported by marine vessels. Schottera nicaeensis a red macroalga Unlikely Usually in ports, under jetties or on jetty piles, or other shaded areas. Known in South Australia from GSV. Ulva fasciata sea lettuce / sea strap lettuce Unlikely Possibly occurs in the GSV and Eyre bioregions. A tropical plant occurring only in the warmer summer months in southern and eastern Australia. Found from intertidal zone to 10m deep, on any firm substrate, in waters ranging from calm to highly wave-exposed. Often associated with high-nutrient areas such as mangroves and bird roosting islands, polluted areas, near freshwater sources, or in ports. 74 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Dinoflagellates Alexandrium catenella toxic dinoflagellate / estuarine dinoflagellate Unlikely Known in South Australia from the Port River area in GSV. Found in temperate to warm temperate coastal and estuarine waters; occasionally reported in tropical waters. Usually found less than 20m, but can also occur in much deeper waters offshore. Cysts found in bloom areas are associated with fine, organic rich estuarine and coastal sediments. The toxins produced by A. catenella are bioaccumulated in fish, molluscs, crustaceans, polychaetes and some echinoderms. Consumers of contaminated organisms suffer from paralytic shellfish poisoning (PSP). There is also some evidence of direct toxicity to fish. Alexandrium minutum toxic dinoflagellate , estuarine dinoflagellate Unlikely Known in South Australia from GSV, Eyre and Murat bioregions; mainly associated with ports, and was introduced by shipping. In eastern GAB, known from Streaky Bay (including cysts in sediment) and Thevenard. Found in warm, temperate, coastal and estuarine waters. Usually found less than 20m, but can also occur in much deeper waters offshore. Cysts in bloom areas are associated with fine, organic, estuarine and coastal sediments. This micro-alga causes serious algal blooms, resulting in water discolouration, loss of oxygen and paralytic shellfish poison (PSP) outbreaks. Causes regular toxic red tides in the Port River in Gulf St Vincent. Alexandrium tamarense toxic dinoflagellate / estuarine dinoflagellate Unlikely Possibly introduced to several parts of southern Australia, including the GSV bioregion, and waters of the southeast, including offshore from Pt MacDonnell. Found inshore (1-20m), in temperate to warm temperate coastal and estuarine waters. Cysts found in bloom areas associated with fine, organic estuarine and coastal sediments. Sometimes associated with shellfish and fish aquaculture, and blooms can cause severe impacts through PSP (paralytic shellfish poisoning) toxins accumulating up the food chain, from zooplankton to shellfish, birds, fish and mammals. 75 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes toxic dinoflagellate / chain-forming dinoflagellate / estuarine dinoflagellate / naked dinoflagellate Unlikely In South Australia, known from Port Lincoln in Spencer Gulf and possible introductions in the GSV and Eyre bioregions. Found in bays and estuaries. Vegetative cells can be distributed throughout the water column during a bloom, and cysts are found in sediments. Toxins produced by G. catenatum can cause Paralytic Shellfish Poisoning (PSP). Toxins accumulated in shellfish (oysters, mussels and scallops) then become toxic to consumers, including humans. Knotted Thread Hydroid Possibly Recorded in South Australia and prefers open waters and good current flow. Grows on variety of substrates, including red algae, seagrasses, ascidians, sponges, bryozoans, shells, stones, crustaceans. Cordylophora caspia a hydroid Very Unlikely Known in Australia from ports in eastern and south-eastern States. May also be present in GSV. Forms small, branched colonies on rock and other firm substrates in shallow brackish to fresh waters, including river mouths. Can withstand a high degree of eutrophication and is present in areas with high levels of run-off and pollution. Halecium delicatulum a hydroid Unlikely Known from estuaries and harbours, mainly in south-eastern Australia. Possibly introduced in the Coorong and Eyre bioregions in South Australia Lives on surfaces of soft-textured sponges and other invertebrates, sometimes bryozoans, occasionally macroalgae. Broad depth range (shallow subtidal to 1,000m). Frequently found in shaded locations such as jetty piles as a fouling species, but no other recorded impacts. Plumularia setacea a hydroid Unlikely Known in Australia from Tasmania, and possibly introduced to GSV in SA, and also other parts of Australia, including south-western WA. Associated with calm coastal waters, typically attached to hard substrates and organisms including other hydroids. Gymnodinium catenatum Hydroids Antennella secundaria (c) J. Watson, Marine Science & Ecology Pty Ltd , at NIMPIS: http://adl.brs.gov.au/marinepe sts/index.cfm?fa=main.spDetai lsDB&sp=6000005529 76 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Jellyfish Cassiopea sp. Upsidedown jellyfish Unlikely Reported in South Australia from GSV bioregion (e.g. Angas Inlet and Port River). Known Mainly from low energy waters in the tropics. Bugula flabellata bryozoan / sea moss Unlikely Amongst other parts of South Australia, has been recorded in the Eyre bioregion, to the east of AW NRM region, but is mainly associated with harbours. It is a major fouling bryozoan in ports and harbours, particularly on vessel hulls, pilings and pontoons. Also reported from off shore oil platforms. Quite often it is found growing with other erect bryozoan species such as B. neritina or growing on encrusting bryozoans. Bugula neritina bryozoan Unlikely Amongst other parts of South Australia, has been recorded in the Eyre bioregion, to the east of AW NRM region, but is mainly associated with harbours. Found on jetty piles, vessel hulls, ships‘ intake pipes and condenser chambers, buoys and similar submerged surfaces In Australia, it occurs primarily on artificial substrates. Cryptosula pallasiana bryozoan Possibly Known in South Australia from GSV and Eyre bioregions (east of AW NRM) and also south-western WA (west of AW NRM). A common fouling organism on seagrasses, drift algae, oyster reefs, artificial structures such as piers and breakwaters, human-made debris, rock, shells, ascidians, glass and vessel hulls. Reported from shallows to 35m deep. Reportedly transported by vessels, and by fishing and aquaculture activity, but colonies found to date in Australia generally do not reach a large size or cover large areas of substrate. Bryozoans (c) California Academy of Sciences 77 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Membranipora membranacea encrusting bryozoan / sea lace Possibly Very widespread globally. Distribution in South Australia uncertain, but does occur here (Bock, 1982). Attaches to kelp and other organisms. Plants fouled by this bryozoan suffer greater blade loss than unfouled plants, probably because fouled blades are fragile and break off easily, and also because fish bite off chunks of blade while foraging on the attached bryozoans (Dixon et al., 1981). Studies in Nova Scotia have shown that changes in cover of this bryozoan on kelp, and in the cover of kelp on the seabed, are reciprocal and seasonal; inter-annual variation in peak cover of this bryozoan on kelp is temperaturerelated; annual decreases in kelp cover and blade size are related to the (c) United States Geological Survey, at Wikimedia Commons degree of infestation by M. membranacea, and not to wave action alone. Severe outbreaks of this bryozoan can occur every few years, and recurrent seasonal outbreaks can have a devastating effect on native kelp populations, which, in turn can aid in establishment and growth of nuisance macroalgae (Scheibling and Gagnon, 2009). Schizoporella unicornis lace coral Possibly Has been reported in the Eyre Bioregion, which includes eastern Great Australian Bight, east of the AW NRM region. Also known from south-western Australia. Can encrust a broad range of substrates, including shell, stone and kelp holdfasts and often forms broad encrustations under boulders or sheltered overhangs. lace coral Possibly Amongst others parts of S.A., has been reported in the Eyre marine bioregion, which includes eastern Great Australian Bight, east of the AW NRM region. An abundant fouling organism that is resistant to anti-fouling paints; can spread rapidly on vessel hulls and provide an area for other species to settle upon. Also attaches to rocks, macroalgae and wooden structures. (c) L. and L. Langstroth, at United States Geological Survey Watersipora arcuata (c) K. Gowlett-Holmes 78 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Zoobotryon verticillatum bryozoan Unlikely Recorded in the GSV and Spencer Gulf bioregions, and in southern WA. Attaches to boats, structures and cooling water intake pipes. A deciduous species, with seasonally growing branches. Masses of Z. verticillatum and drift algae frequently become entangled around submerged hard surfaces, and may significantly impact the settlement and fitness of hard-bottom sessile invertebrates (Walters and Abgrall, 2000, cited by Hayes et al., 2005). Unlikely Occurs as one of the fouling organisms in harbours and bays around the world. It fouls many living (worm tubes, mussel shells, encrusting bryozoans, etc.) and non-living (wood, bark, styrofoam floats etc.) substrates. Known in S.A. from Gulf St Vincent and associated mainly with harbours. pileworm / sedentary worm Unlikely Known in southern Australia from ports in WA, Victoria and NSW, and possibly occurs in SA in GSV. In Australia, associated with soft sediment bottoms in ports and estuaries, but on a global scale not restricted to such areas. May be transported by vessels, and also by natural means. Boccardia proboscidea Californian polydorid / spionid worm Unlikely Amongst other parts of South Australia, has been recorded in the Eyre bioregion, to the east of AW NRM region. Burrows into rocks, sediments and shells of oysters and other bivalves, and increases risk of parasitism in bivalves. Survives in a variety of habitats but mostly recorded in nutrient-enriched areas, such as sewage outfalls. Transferred by shipping, fisheries activity and natural dispersal. Euchone limnicola fanworm Very Unlikely Occurs in soft sediments of harbours, and no confirmation of presence in South Australia, but might be present in lower South East - Otway Bioregion. Nodding Heads (Kamptozoans) Barentsia benedeni nodding head Annelid Worms (including Polychaetes) Alitta succinea (formerly Neathes succinea) 79 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Hydroides elegans Fouling Serpulid Unlikely Point of origin uncertain. Recorded in South Australia in the Gulf St Vincent and Eyre bioregions, but also occurs in southwestern Australia. Known mainly in harbours in the tropics and sub-tropics. Occurs on both natural and artificial substrates. Highly tolerant of polluted waters. Can grow in high densities, heavily fouling any new immersed structure - natural or artificial. It creates microhabitat for other species, changes ecosystem dynamics and competes with other species for food and space. The fouling habit of this species also causes problems for seawater piping systems. Myxicola infundibulum Slime Featherduster Worm Possibly Known in Australia from Tasmania, and Gulf St Vincent in SA. In SA, possibly also occurs in Eyre bioregion. Lives in coarse, clean sands, in which it secretes a gelatinous tube. Usually found from 1-50 m, but can occur as deep as 500m. It commonly inhabits areas where the temperatures range from 2-17 °C. Transported by vessels. No recorded impacts. Polydora ciliata / Polydora websteri Bristleworm / English Polydorid Unlikely To date, known in South Australia from Gulf St Vincent, in disturbed areas. Found in crevices within rocks, on soft sediments and is also recorded as a boring species - burrowing into rocks and mollusc shells. A serious pest for oysters and mussels, as it burrows into shells, weakens them and makes shells prone to predation by organisms such as crabs. It can also alter habitats by increasing the amount of mud in an area (due to burrowing activities). Pseudopolydora paucibranchiata Japanese Polydorid / Elkhorn Slough Spionid Unlikely Has been reported in the Eyre marine bioregion, which includes eastern Great Australian Bight, east of the AW NRM region. Probably spread by ballast water, hull fouling and Japanese oysters and can become a dominant member of the infaunal community in infested areas. However, more likely to occur in disturbed areas, and in low energy environments (sheltered bays, mangroves etc. (c) J. Finn, Museum Victoria, at Pests and Diseases Image Library: http://www.padil.gov.au 80 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Sabella spallanzanii European Fan Worm / Sabellid Unlikely Known in S.A. from the eastern side of GSV bioregion, particularly North Haven, West Lakes, Port River, and Glenelg. Found mainly in shallow subtidal areas from 1 to 30m deep, especially harbours and wavesheltered bays. Colonises both hard and soft substrates, often anchored to hard surfaces within soft sediments, or on wharf piles and facings, channel markers, marina piles and pontoons, and submerged wrecks. Can compete with razor fish, can displace scallop and mussel species, and spreads very rapidly. Crassostrea gigas Pacific Oyster Unlikely Known in S.A. from Murat, Eyre, Spencer Gulf, Northern Spencer Gulf and GSV bioregions. Found mainly adjacent to oyster leases. In eastern GAB, known from Murat Bay, Streaky Bay, and Smoky Bay. Attaches to almost any hard surface (especially rocks) in shallow sheltered waters such as estuaries, but also found in muddy or sandy areas. Oysters will also settle on adult oysters of the same or other species. Although recorded in a Ceduna area, east of AW NRM region, lack of estuarine habitat and lack of aquaculture facilities may impede settlement in the Eucla Bioregion (i.e. AW NRM waters). Maoricolpus roseus New Zealand Screw Shell Possibly According to DEH SA (2001), it has been recorded from the ―west coast of South Australia‖, but NIMPIS (2010) and other references report it only from south-eastern Australia. Introduced to Tasmania in the th early 20 century, and now occurs also in Victoria and NSW. Very abundant in Bass Strait. Has potential to spread westwards. Appears to prefer areas of coarse or firm substrate, moderate to strong currents, and can tolerate broad range of conditions (depths of 1m -130m and temperatures between 8-20 degrees C). The impacts of dense populations of M. roseus on native species have yet to be determined, but M. roseus could possibly reduce numbers of native gastropods and bivalves via competition for food and space. Transferred naturally, or by vessels (including ballast water) or aquaculture (e.g. attached to oysters). Molluscs (c) T. Alexander, at http://www.ausmarinverts.net 81 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Musculista senhousia Asian Date Mussel / Asian Bag Mussel Unlikely Known in South Australia from GSV. Occurs in soft sediments and also wharf piles and other human-made structures in harbours. Mytilus galloprovincialis Mediterranean Blue Mussel Possibly May be native, but origin uncertain. Possible that this species came to Australia on hulls of early sailing ships when the continent was colonised by Europeans. Farmed in South Australia. The taxonomy is uncertain. This species has been known under several names in Australia, including Mytilus edulis, M. planulatus, and M. edulis planulatus. There is some question as to the valid name, although recent genetic studies indicate that the Australian form is closer to M. galloprovincialis than any other (Beu, 2004, cited by CSIRO, 2009). May hybridize with other Mytilus species. Common in harbours on jetty piles and exposed rocks, and is found from the intertidal to about 15m deep. (a beach clam from New Zealand) Possibly Reported to Academy of Natural Sciences 2006 to occur in the Great Australian Bight. Very little information. Not listed in Hayes et al. 2005 or NIMPIS 2010. New Zealand Greenlip Mussel Unlikely Previously recorded in GSV bioregion: P. canaliculus was successfully eradicated from shipping channel adjacent to Outer Harbour wharf in 1996 (PIRSA 1999, cited by McEnnulty et al., 2001). However, according to the PIRSA Aquaculture Public Register in August 2003, a licence was provided to farm this species, amongst others, at Rabbit Island near Pt Lincoln. (c) CSIRO Marine and Atmospheric Research Paphies (Mesodesma) ventricosa (c) New Zealand Ministry of Fisheries Perna canaliculus 82 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Polycera hedgpethi Hedgpeth‘s Dorid Unlikely To date, known in South Australia from GSV, in disturbed areas, such as Port River and North Haven. Ruditapes largillierti = Venerupis largillierti Venus clam Unlikely Known in Australia from Tasmania, and GSV in South Australia. Found in fine, muddy to clean sand or fine gravel or seagrass beds, from low intertidal to ~ 20m in depth, usually on wave-protected shores, harbours and estuaries. Theora lubrica East Asian Bivalve / Asian Semele Very Unlikely Recorded in South Australia from northern Spencer Gulf. Associated with ports. Teredo navalis Naval Shipworm / Common Shipworm / Atlantic Shipworm Possibly Generalised distribution across southern Australia includes Eucla Bioregion, of which AW NRM marine region is part. Bores into wooden structures such as boats and harbour piles. European Clam / Atlantic Mediterranean Bivalve / Common Basket-Shell Very Unlikely Known to date in S.A. only from the South East / Victorian border. Mainly in thick muddy sand in polluted environments. Has fast growth rate and high tolerance of many environmental conditions, including low oxygen levels. Can achieve very high population densities and therefore has potential to compete with native species for food and space, including commercial species such as scallops, possibly affecting their recruitment. (c) The Full Wiki : http://www.thefullwiki.org/Shipworm Varicorbula gibba 83 Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Zeacumantus subcarinatus New Zealand Creeper Shell / mud snail Possibly Introduced to Australia from New Zealand during the 1930s, and now broadly distributed across the southern coast. There are records from east of the AW NRM region – e.g. Streaky Bay. Carcinus maenas European Shore Crab / European Green Crab Unlikely Known in SA from GSV and Coorong bioregions, including Barker Inlet, Outer Harbor, West Lakes, Port Stanvac, Port Noarlunga, Fleurieu Peninsula and the Coorong. Associated with soft sediment bottoms in ports and estuaries. Eurylana arcuata cirolanid isopod / sea slater Very Unlikely Introduced from New Zealand, and known in Australia from New South Wales and the South Australian gulfs (including Port Noarlunga and Port Willunga in Gulf St Vincent). Megabalanus tintinnabulum Acorn Barnacle Possibly Very widespread in Australia. No published records from GAB (and none in S.A., according to NIMPIS 2010), but occurs in bioregions east and west of that region, and in most other parts of Australia. A common fouling species, it is found in many types of habitat, such as rock and boulder areas, pylons, wharves, vessel hulls and other organisms such as mussels and algae. Main impacts relate to fouling. Occurs from intertidal to about 40m depth and can withstand water temperatures up to 35 °C. Photo of a related species: Zeacumantus lutulentus. (c) G. Bould, at Wikimedia Commons Crustaceans (c) Henry & McLaughlin, 1986, at NIMPIS: http://adl.brs.gov.au/marinepests/inde x.cfm?fa=main.spDetailsDB&sp=600 0009302 84 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Monocorophium acherusicum Mediterranean Corophiid (an amphipod) Unlikely Known in South Australia from GSV. Occurs subtidally on sediments or where silt and detritus accumulate among fouling communities such as algae, ascidians and bryozoans, and human-made installations e.g. wharf piles, rafts and buoys. Constructs conspicuous, fragile Ushaped tubes of silk, mud and sand particles. Reportedly transported by fisheries and aquaculture activity and vessels. Monocorophium insidiosum English Corophiid / amphipod Very Unlikely Known in Australia from southeastern States, and from Perth area in WA. May also occur in GSV. An estuarine species, and often associated with mud, algae, seagrasses and hydroids in ports, where it builds mud tubes. It is common in areas of high turbidity, especially in areas with high organic pollution levels. Paracerceis sculpta Sponge Isopod Unlikely Known in South Australia from GSV. Occurs intertidally in waveexposed areas. Mainly associated with macroalgae, in sponges and under stones. Also found amongst the fouling community on buoys, vessel hulls and other humanmade structures. Reportedly transported by vessels, but there are no recorded impacts. solitary ascidian Unlikely To date, known in South Australia from GSV and Eyre Peninsula. In Australia, found mostly in harbours and other protected embayments. Ranges from intertidal waters to 50m depth, and attaches to clay, stones, rocks, algae and wharf piles, where it can be the dominant fouling species. Reportedly transported by vessels, fishing and aquaculture activity. Ascidians Ascidiella aspersa 85 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Botrylloides leachi colonial ascidian / Dark Ascidian Possibly Possibly introduced into most parts of the Australian coast, including Murat bioregion (adjacent to Eucla, which includes AW NRM waters), Eyre, Spencer Gulf, Northern Spencer Gulf and GSV bioregions in South Australia. Has been recorded at Topgallant lsland in the eastern Great Australian Bight. An encrusting species from the lower intertidal and shallow subtidal, growing on stones, shells (including oysters), macroalgae, seagrasses, other fouling species; also human-made structures such as floats, ropes and aquaculture cages. Nudibranchs, gastropods and flatworms feed on this colonial ascidian. Can be a dominant competitor, overgrowing and excluding many other fouling species. Transferred by vessels, fishing and aquaculture activity. Star Ascidian Possibly Known in South Australia from the GSV and Eyre bioregions (including southern Eyre Peninsula), and also from southwestern WA. Attaches to rocks, macroalgae, seagrass, other fouling organisms, floats, wharf pylons, vessel hulls and other submerged human-made structures. Colonies of B. schlosseri are eaten by crabs, gastropods, nudibranchs and flatworms. Colonies can occur from the very shallow subtidal to almost 200m depth. Can be transferred by vessels, fisheries activity and aquaculture. solitary ascidian Unlikely Known from port areas in several parts of southern Australia, including Port Adelaide and Outer Harbour area in GSV bioregion in SA. Commonly found in dense aggregations on rocks, algal holdfasts, seagrass, shells and artificial structures such as pylons, buoys and ships‘ hulls. It is found in the low intertidal and shallow subtidal zones of enclosed and semi-protected marine embayments and estuaries. Australian populations appear to be in decline in port areas, but it is noted that this species is reported to be easily transferred by vessels, fishing activity and aquaculture. (c) K. Gowlett-Holmes, in Stafford and Willan (2007) Botryllus schlosseri (c) K. Gowlett-Holmes at NIMPIS: http://adl.brs.gov.au/marinepests/ind ex.cfm?fa=main.spDetailsDB&sp=60 00006015 Ciona intestinalis 86 Table 7 (continued): Marine pests reported in South Australia. Latin Name Common Name Likelihood of Presence in AW NRM Region Supporting Notes Styela clava Leathery Sea Squirt Very Unlikely Reported by NIMPIS 2009a to occur in South Australia, but Hayes et al. 2005 and Kott (2006 in ABRS 2009) reported that this species is known to date only in Victoria. Known to foul vessels, aquaculture and fishing equipment and other artificial structures, and may also be translocated with oyster spat and oyster transfers. Styela plicata solitary ascidian Unlikely Reported by NIMPIS 2002 to occur in South Australia gulfs and eastern GAB, but not as far west as AW NRM region. Associated with ports and harbours and aquaculture. Research Disturbance to Australian sea lions from researchers undertaking work within colonies is possible unless research activities are conducted appropriately (Director of National Parks, 2005). Benthic research using sleds, dredges and cores can damage the sea floor, destroy benthic flora and fauna in the dredged or cored area, increase turbidity, and degrade the habitat quality for mobile organisms that utilise the structure and composition of the benthic cover for feeding, reproduction, living, or shelter from wave action / storms or predators. This is particularly so in areas where slow-growing species form a significant part of the benthic cover, such as sponges and bryozoans. Land-based and Coastal Discharges The broad shelf of the GAB is largely uninfluenced by run-off from the land, because there are no major or minor rivers draining into the Bight (DEWR, 2006). Due to this absence of river runoff, coupled with the very low population density, absence of large urban centres, absence of coastal industry, and absence of agriculture, major land-based discharges are not an issue in the region. However, at a local scale, the proliferation of tracks in some parts of the cliff top and cliff face within the AW NRM region, have resulted in gully erosion (Caton et al., 2007). Cliff erosion in this area is a natural process, but the nearshore marine impacts of accelerated erosion from human-made tracks (and consequent increased rate and quantity of cliff materials deposition into the shallow marine environment) has not been investigated. 87 In parts of the AW NRM, the increased number of tracks through dunes during the past two decades has also caused some dune destabilisation and erosion, and this is discussed in more detail in Caton et al. (2007). Mining, Minerals, Oil and Gas Exploration Figures in section 2.7 show the parts of the Great Australian Bight in which oil and/or gas exploration permits and exploration wells exist, and areas where seismic surveys have been undertaken. None of these activities have occurred in the shallow waters of the AW NRM Region. However, it is noted that significant impacts can occur from such activities in deeper waters (managed by the Commonwealth). No demersal (sea floor) trawling is allowed in the allowed in the Commonwealth-managed Benthic Protection Zone of the Great Australian Bight Marine Park (see Table CC in 3.3 Notes on Current Protection and Management); however, such activity is permitted outside the benthic protection ―strip‖. Also, within the area of the GAB Marine Park Commonwealth waters, petroleum exploration leases have been approved, and petroleum exploration in the area has involved seismic testing as well as proposals for core sampling (i.e. drilling into the ocean floor). Scientific exploration for minerals in the area is also a disruptive process, which includes dredging and sediment core sampling. Impacts from benthic drilling include destruction of benthic fauna; siltation, and damage to colonies of filter-feeding organisms (such as sponges); erosion of bottom sediments when benthic cover is removed; provision of conditions of opportunistic organisms to proliferate (thereby altering community structure, including species composition and abundance); and noise impacts (and disturbance to various species that may use the area for feeding, resting etc). Offshore seismic surveys involve the use of high energy noise sources operated in the water column to probe below the seafloor. Almost all routinely used seismic sources involve the rapid release of compressed air to produce an impulsive signal. These signals are directed downwards through the seabed, and then reflected upwards again by density or velocity discontinuities within the underlying rock strata. The technique is commonly used in oil and gas exploration (McCauley et al., 2000). There is a large and increasing literature on the impacts of seismic surveying on whales, dolphins, sea turtles, and fishes. In 2003, Pidcock et al reported on the potential sensitivity of marine mammals to mining and exploration in the Great Australian Bight Marine Park Marine Mammal Protection Zone. Despite a significant number of publications on this topic, it is noted that the general conclusion of the authors was that ―there is still insufficient definitive data to determine, with any degree of certainty, what is likely to occur should petroleum operations be allowed in the Marine Mammal Protection Zone under (a future) management plan”. Impacts on marine mammals from the gas and petroleum industry can include the following (from Pidcock et al., 2003):   noise from seismic survey, drilling activities, boat or rig operations; collisions with ships involved with petroleum industry activities; 88   chemical impacts associated with oil spills or other sources of industry-based pollution; and indirect impacts through effects on prey species populations from industry activities such as seismic surveys or events such as oil spills. It is generally agreed by those that study effects of mining, minerals, oil and gas exploration industries on marine mammals in Australian waters, that the primary potential source of impact or disturbance from the industry is through sound (Pidcock et al., 2003). One example is the work of McCauley and colleagues during the mid to late 1990s in Western Australia, who undertook at 3-year survey for the Australian Petroleum Production and Exploration Association (APPEA) in conjunction with the Energy Research and Development Corporation (ERDC). Results of the project showed the following, in response to seismic vessels and associated air gun signals (from McCauley et al., 2000, and McCauley 2004):       migrating humpback whales made avoidance manoeuvres in response to a 3D seismic vessel at approx. 2 to 4 km, then allowed the seismic vessel to pass no closer than 3km. Humpback pods containing cows which were involved in resting behaviour in key habitats were more sensitive and showed an avoidance response estimated at 7-12km from a large seismic source. (Generally, whales are more susceptible to noise while resting or breeding compared to during migration or feeding, but resting is a particularly important behavioural state for cow-calf pods); male humpbacks were attracted to a single operating air gun due to what was believed the similarity of an air gun signal and a whale breaching event. based on the response of captive animals to an approaching single air gun and scaling these results, sea turtles displayed a general 'alarm' response at an estimated 2km range from an operating seismic vessel and behaviour indicative of avoidance estimated at 1km. captive fishes showed a generic fish 'alarm' response of swimming faster, swimming to the bottom, tightening school structure, or all three, at an estimated 2-5 km from a seismic source. Also, captive fish exposed to short range air gun signals were seen to have some damaged hearing structures, and captive squid showed a strong startle response to nearby air gun start up and evidence that they would significantly alter their behaviour at an estimated 2-5 km from an approaching large seismic source (McCauley et al., 2000). Although this research project did not take place in the GAB, the results may be relevant, because the affected faunal types all occur in the GAB either permanently or periodically, and may display a similar response to seismic testing. According to a literature summary cited by Pidcock et al. (2003), for seismic noise, most whales show avoidance behaviour at 180 dB. Importantly, the greatest risks to marine mammals are considered to be posed by exploration activities (compared with production activities). 89 A number of exploration companies in southern and western Australia have undertaken studies of seismic impacts on whales, as a requirement under the Commonwealth‘s guidelines on the application of the EPBC Act to interactions between offshore seismic operations and cetaceans. In some cases, shut-down procedures have been instigated during operations when whales were seen within a 5km and 3km mitigation zones (e.g. ROC Oil 2002 and 2003 programs, cited by APPEA, undated). Ship strike is another potential impact for whales in Commonwealth-managed waters, but far less so in State waters at the Head of the Bight, where few vessels are permitted. Ship strike is generally considered to be a serious risk for Southern Right Whales (summary in Pidcock et al., 2003) particularly young right whales which tend to rest on or near the water‘s surface. Climate Change Changes in ocean temperature, currents, levels of ocean stratification, wind strength, rainfall, sea level, surface irradiance, ocean chemistry and extreme weather conditions are some of the key aspects of variables climate change impacts in marine environments (Booth et al., 2009). Climate change requires global effort for mitigation, and CO2 emissions are unlikely to cease this century despite an increase in policies to mitigate climate change by nations around the globe (Babcock et al., in Poloczanska et al., 2008). However, in a ranking of vulnerability of marine life to climate change impacts within Australia‘s 7 large marine regions, the southwestern part of Australia (which includes the Great Australian Bight), was considered the least vulnerable to change (Hobday et al., 2006), compared with the central eastern and southeastern parts of Australia (the most vulnerable). The assessment included 29 indicators in five dimensions (biological, regional, climate change, fishing stress and other anthropogenic stressors) (Hobday et al., 2006). In the eastern Great Australian Bight, cooling due to increased winds and upwelling is predicted to counteract the influences of ocean warming to some extent (Hobday et al., 2006). Compared with south-western Australia and south-eastern Australia, there has been little investigation of the mechanisms or likely impacts of climate change specifically in the Great Australian Bight. In south-eastern Australia, sea surface temperatures (SST) have been shown to have risen ~2°C since 1925, with inter-annual variation in extremes also increasing (Figueira and Booth, 2009, cited by Booth et al., 2009). Warming may have occurred in Western Australia, but the data are less clear. Given that the Leeuwin Current from Western Australia is a significant seasonal influence in the Bight, it is highly likely that changes in this current over time that can be attributed to the influence of global warming, will have an impact on the ecosystems of the GAB. Climate change is predicted to enhance the strength, annual duration and southward extent of boundary currents such as the Leeuwin Current, and its associated eddies (Ridgeway, 2007, cited by Booth et al., 2009). Furthermore, a gradual increase in sea surface temperature (SST) has been observed in south-western Australia, and this may be linked to air-sea heat flux into the southern Indian Ocean (Pearce and Feng, 2007; Caputi et al., 2009, cited by Booth et al., 2009). Possible impacts of climate change on several key temperate marine habitats and groups are discussed below. 90 Temperate Rocky Reefs: Rocky reefs have been ranked internationally as one of the five most threatened marine ecosystems (Halpern et al., 2007). Rocky reefs, commonly dominated by brown or mixed red/green/brown macroalgae cover in shallow waters and attached invertebrates and red macroalgae in deeper waters, form a large component of the AW NRM marine region, and increasing sea temperature is considered to be a significant (and increasingly serious) stressor on temperate rocky reefs, due to the effects of global warming. There is concern that the distribution and abundance of macroalgae on temperate reefs in southern Australia may be significantly affected by global climate change, because climatic stability over geological time has been one of the key conditions underpinning the evolution of the marine flora in the southern region (Wernberg et al., 2009). Some changes may already have occurred, but these relate to the well publicised decline of giant kelp Macrocystis pyrifera in Tasmania (for which the causes have still not been unequivocally determined), and the southward shift of 3 species of canopy-forming macroalgae on the east coast of Australia (Wernberg et al., 2009). Apart from work on eastern Tasmania reefs (Crawford et al., 2000; Edgar et al., 2005; Ling, 2008; Ling and Johnson, 2009; Ling et al., 2008, 2009), there are very few examples from Australia of climate change impacts on rocky reefs in Australia, because of a lack of long-term and broad-scale data sets. The Tasmanian work has indicated that current and increasing water temperatures, and movement further south of the warmer Eastern Australian current, have caused a southerly range shift for sea urchins from eastern Australia, leading to overgrazing and loss of algal habitat. Ling (2008) predicted that although grazed barrens are relatively uncommon in southern Australia, increased abundance and range (similar to that observed in Tasmania), could occur as a result of ocean warming. This could have very serious consequences for the high proportion of southern Australian endemic macroalgae and the associated fauna, as poleward range retreat is prevented by a lack of contiguous southern coastal habitat (Ling, 2008). The range of a number of other invertebrate species from eastern Australia has also increased southwards during the past 50 years (Poloczanska et al., 2008). It is predicted that in future, rising temperatures will result in changes in species distribution, abundance, and patterns of reproduction (e.g. frequency and timing of spore release, and changes in reproductive output), including further range-shifts of both macroalgae and invertebrates on rocky reefs, with local extinction of species that have northern range limits along the southern coastline of Australia. Such changes are predicted to happen progressively through 2030 to 2100 (Wernberg et al., 2009). This may have important consequences for community composition on rocky reefs, if key structural and functional species such as brown canopy-forming macroalgae are adversely affected by increasing ocean temperature. A summary of likely changes to kelp-dominated habitats includes (i) a shift in the distributions of kelp and other temperate reef species southward, due to increasing temperatures and decreases in nutrients, and such a shift may imperil some species due to the bounded southern coastline; (ii) changes in storminess and physical disturbance regimes will shift the distributions and compositions of kelp ecosystems; (iii) changes in rainfall and watershed integrity will affect kelp habitat; (iv) outbreaks of herbivores might reduce the cover of kelp forest stands in some areas (Hobday et al., 2006). 91 In some areas, the combined effects of climate change and other stresses (such as poor water quality) is predicted to reduce the resilience of temperate reef communities to perturbations such as storms, diseases, and invasive species. Many of these perturbations may increase in frequency and/or severity in response to climate change (Wernberg et al., 2009). However, it is noted that although the flora and fauna in the waters of the AW NRM region may be subject to the increasing and long-term effects of climate change, they are largely not subject to any other stressors (other than potential recreational fishing impacts), due to the remoteness of the region, lack of pollutants from the coast, low population density, and the legislative protection afforded by the Great Australian Bight Marine Park. Seagrasses: Seagrass beds can also be vulnerable to higher temperatures. One example comes from Spencer Gulf in South Australia, where a rapid die-off of 12,717 hectares of seagrass was attributed to the effects of a hotter than usual summer during an El Niño year (Seddon et al., 2000). Sea surface temperature increases of 1-2°C across southern Australia is predicted by 2070 (with greatest warming in SE Australia/Tasman Sea due to strengthening of the East Australian current), an also, an increase in incident solar radiation on the sea surface in Australian waters is expected by 2070, in the order of between 2 and 7 units W m-2 (Hobday et al., 2006). Both the increased temperatures and the increase in solar radiation on the surface could adversely affect shallow marine biota such as intertidal and shallow subtidal seagrasses. Future reduction of seagrass habitat, which forms key nursery area for juveniles of many species, is expected to occur, due to the impacts of climate change (Booth et al., 2009), and seagrass physiology and morphology may also change (Hobday et al., 2006), in response to increased solar radiation, increased CO2, and increased sea surface temperature. Decreasing cover and extent of seagrass may have flow-on effects, such as a reduction in the populations of various temperate coastal fishes. At a national scale, a summary of some of the changes likely in future includes: (i) increases in seagrass biomass as CO2 levels rise and enhancement of the role of seagrass in coastal carbon and nutrient cycling; (ii) shifts in species distribution southwards, and alteration of species composition of seagrass beds as temperatures warm; (iii) alteration of frequency and timing of flower and seed production; (iv) increase in seagrass bed destruction if storm regimes become more frequent or more severe; (v) disappearance of ultraviolet (UV)- intolerant species from shallow waters as UV radiation levels rise (Hobday et al., 2006). Fishes: Changes in distribution (latitudinal changes along eastern and western Australia, and both latitudinal and longitudinal changes along the south coast, including the GAB) may be one of the key impacts of climate change on temperate fishes. Many benthic and demersal fishes may be particularly affected, because of their localised reproduction and reduced ability to disperse, hence current distributions and abundances are highly likely to change over time (Hobday et al., 2006), including localised extinctions of populations. Climate change may affect temperate fishes in many ways, including changes to larval transport / dispersal and population connectivity, settlement and recruitment (early post-settlement survival and over-wintering, particularly in relation to changes in currents), growth rates, assemblage structure and range shifts, spawning and egg production (including nursery grounds) (Booth et al., 2009). Examples of some of the likely impacts of climate change on temperate fishes include the following: (i) contraction of spawning period and southward movement of the spawning area for species which spawn late autumn/winter (e.g. King George Whiting); (ii) east / west range shifts along the South Australian coast, and also possibly in depth distribution, of Australian Pilchard 92 (Sardinops sagax); (iii) changes to strength of the Leeuwin Current and warming SST are likely to enhance settlement and early survival of warmer-water fish species, with the opposite effect for cooler-water southern species; (iv) reduced habitat for cooler water southern species, including commercially important species; (v) establishment of fish species at southern latitudes that were previously absent, e.g. Western Australian coastal fishes may establish/increase abundance in South Australian waters, and (vi) changes in growth rates of some species, because growth warmer waters are often inversely correlated with ocean productivity, and hence fish food resources (Booth et al., 2009). One of the major impacts likely to accelerate in future, is the north-south range shifts of temperate and tropical fish species, due to the effects of warming sea surface temperature on species‘ tolerances, coupled with ocean current changes linked to potential for larval dispersal (Booth et al., 2009). In south-eastern Australia there is some research on early post-recruitment survival of tropical reef fishes, which disperse to southern areas via the SE Australian coast. For these species, an overwintering survival threshold of around 17.5°C has been demonstrated, but this threshold was exceeded in 2001 and 2006 (warmest winters in 150 years: Figueira and Booth, 2009, cited by Booth et al., 2009), and this resulted in significantly greater numbers of several tropical species over-wintering in southern latitudes. Furthermore, in the last decade 34 fish species that were previously absent or rare south of the Bass Strait have either established or increased in abundance further south (Last, CSIRO, pers. comm., cited by Richardson and Poloczanska, in Poloczanska et al., 2008). It is quite possible that such a phenomenon could also occur along the southern coast of Australia in future, for tropical and sub-tropical species from Western Australia, driven southwards by the Leeuwin Current. Little is known of the larval dispersal patterns of most temperate fishes, so impacts of changes in current patterns on dispersal are unclear, but one possible impact may be adversely affected larval fish development and survival due to increased ultraviolet radiation (Hobday et al., 2006). In south-western Australia and through the Great Australian Bight, it is predicted that with increased warming effects, future changes in the strength and timing of the Leeuwin Current may have major implications for the distribution and abundance of some species, including West Australian Salmon and Australian Herring in waters off South Australia (Hobday et al., 2009). Changes in productivity, for example due to increased coastal upwelling, may lead to increases in abundance of some species, particularly of small coastal pelagic fishes, such as sardines and anchovy, in the large upwelling system between Cape Otway in Victoria and the central Great Australian Bight (Hobday et al., 2009). However, due to limited evidence for observed changes (which makes prediction difficult), coupled with absence of information on species habitat tolerances and empirical models for future prediction of species ranges and abundances, the confidence in predictions is low to medium (Hobday et al., 2009). In summary, it is possible that significant range shifts – including more tropical species in southern waters; southern species shifting south, driven by sea surface temperature warming; change or strengthening of boundary currents, habitat losses and gains – will be strongly evidenced by 2030, with possible large-scale changes in fish assemblage structure and fishery outputs by 2100 (Booth et al., 2009). 93 Seabirds: At an Australia-wide scale, oceanic warming is expected to shift the distribution of some seabirds southwards, and there is already some evidence of distributional changes in tropical Australian seabird populations (Hobday et al., 2006). Other possible changes include altered reproductive success; earlier nesting and laying; loss of nesting or feeding habitats; and changes to migration and foraging patterns, due to the effects of changing current patters on productivity and food sources etc (Hobday et al., 2006). Soft Sediment Fauna: A summary of some of the changes likely in future includes: (i) sensitivity of populations of sediment fauna warming, decreases in ocean pH, changes in primary production, hydrology, storminess and extreme weather events; (ii) a shift in the ranges and distributions of many species of soft sediment fauna, differentially leading to new assemblages of organisms; (iii) acidification (see following paragraph) will erode the shells of molluscs, corals, benthic foraminifera, and other groups with carbonate shells; (iv) changes in water column primary production may alter abundances, community interactions, and benthic-pelagic coupling; (v) changes in prevailing current dynamics may affect soft sediment systems considerably, and (vi) changes in freshwater input may modify the benthic community over vast areas (NB the last point of least relevance to AW NRM marine region, which has very little freshwater input). Oceanic acidification, and also changes in storm patterns, may also result from climate change. Although there are currently no examples of such impacts for southern Australian ecosystems (Wernberg et al., 2009), the predictions are serious. For example, McNeil and Matear (2008) presented an observational and empirical analysis of the seasonal variability in pH and carbonate ions in the Southern Ocean. Their work indicated strong variation in the seasonal amplitudes of pH and carbonate, with an intense minimum of carbonate ions during the wintertime. This coincides with the development of some prominent calcifying plankton species. Such species are predicated to experience detrimental ocean acidification conditions ~30 years earlier than previously thought, with a predicted threshold of 450ppm carbon dioxide for the Southern Ocean wintertime carbonate unsaturation projected for the year 2030 (depending on future emission scenarios) (McNeil and Matear, 2008). In the future, it is likely that ocean acidification will increase (possibly in the order of a 0.2 unit decrease in pH: Hobday et al., 2006, or a 0.3-0.4 unit decrease by 2100: Wootton et al. 2008, cited by Wernberg et al., 2009), with negative impacts on calcareous macroalgae and other calcifying organisms. Examples include corals, calcareous plankton (e.g. coccolithophores), foraminifera, and larger fauna with calcareous shells, such as echinoderms, bivalve and gastropod molluscs, and larger crustaceans which strengthen their external cuticle with calcium carbonate deposits. The changes in the ocean‘s chemistry (i.e. decrease in the pH caused by oceanic uptake of anthropogenic carbon dioxide), will reduce the capacity for such organisms to secrete calcium carbonate (reduced rates of calcification), and /or will increase the rate of dissolution; i.e. the concentration of carbonate ions will be lowered to a point where the shells of many marine organisms will begin to dissolve. Many invertebrate shells are constructed of calcium carbonate, and this is also the structural support for the plant body of calcareous algae. On the temperate coast of southern Australia, calcareous macroalgae crusts occupy significant areas of hard substrate (up to 80% in some exposed areas), dominating space beneath canopies (Wernberg et al., 2009). Recent experimental work in Australia has shown that acidification associated with conservative projections of future CO2 concentrations (550 ppm) will have negative effects on the growth and recruitment of coralline algae (Russell et al., 2009, 94 cited by Wernberg et al., 2009). The work of Russell et al. (2009) also indicated that the effects of increase CO2 may be worse in nutrient-enrichment areas, where opportunistic turfing algae may proliferate in place of calcareous algae. Work in the northern hemisphere has shown that increased CO2 and temperature have greater negative effects in combination than in isolation (Martin and Gattuso, 2009, cited by Wernberg et al., 2009). Decalcification and reduced rates of calcification may be particularly important future impacts in parts of the Great Australian Bight shelf, where calcifying organisms are abundant. Ocean warming may also influence the spread of pest species. For example, Glasby and Gibson showed that both the tropical native invasive forms of Caulerpa taxifolia grew up to ten times faster in warmer waters (22–25°C) compared with cooler water (15–18°C), in terms of biomass, and length and number of both fronds and stolons. Given avoidance by certain fish species of C. taxifolia, and its encroachment into and displacement of preferred seagrass beds, potential indirect effects of sea surface temperature rise are likely to affect seagrass-associated fish populations in shallow waters. However, it is noted that Caulerpa taxifolia may have a low probability of invading remote and high energy areas such as AW NRM region (see section on Marine Pests). Another consequence of global warming may be the increased incidence in future of harmful algal blooms of phytoplankton groups that can proliferate due to thermal stratification and acidification of the ocean. Warmer temperatures may also result in expanded ranges of some harmful species, such as tropical marine dinoflagellates (Moore et al., 2008). Two other habitats that are vulnerable to impacts from increasing sea temperature (hard shelf and epi-pelagic ecosystems) are major components of the waters further offshore from the region, but the discussion here will be limited to inshore habitats, which form part of the AW NRM region. Ranking of Threats and Proposed Actions Table 8 below summarises the main issues discussed above, and lists proposed actions to mitigate threats and threatening processes. A number of recommendations apply to deeper waters seaward of AW NRM region, and these are included with respect to the ecological connectivity between shallow and deep waters of the region. For example, some of the fishes and sharks caught commercially in deeper waters seaward of the NRM, have nursery areas in shallow waters. Also, sea lions that occur in the AW NRM region feed in offshore waters, where they may be subject to bycatch mortality, or mortality or injury due to marine debris. 95 Table 8: Summary of issues / threat categories, area of concern (within AW NRM and/or seaward of AW NRM region), proposed actions, and priority ranking for actions. Issue / Threat Category (and Area of Concern) Proposed Action Priority of Action Fishing (AW NRM) A program is required to educate and inform fishers and the general public about the conservation status of species at risk, such as mulloway, western blue groper, southern bluefin tuna, school shark and the whaler sharks. These species should not be fished in the Great Australian Bight Marine Park if possible (see below), including waters of the AW NRM, due to their vulnerable population dynamics. High Fishing (AW NRM) It is recognised that mulloway fishing is a key recreational activity at the Head of the Bight, and a if managers and other decision-makers determine that the social and economic consequences of prohibiting mulloway fishing are too great, then at least measures (such as hook specifications, restrictions on time permitted to ―play‖ a fish‖, and minimum handling etc) should be undertaken to try to ensure that mulloway that are caught and released suffer minimum post-release mortality. Some examples of ways in which mortality can be reduced include (i) for mulloway which have swallowed the hook: cutting the line rather than trying to remove the ingested hook (NB fish which are mouth-hooked suffer much lower mortality rates, regardless of whether the hook is removed or the line is cut), and (ii) cutting the line underwater (and thus reducing the exposure of the fish to air). High Fishing (AW NRM) Determine relative numbers of mulloway in the GAB, compared with other parts of the State, and monitor catches in the area over space and time. High Fishing (AW NRM) Separate quotas for greenlip and blacklip abalone in Region B of the Western Zone , to assist monitoring of stock status. Medium Fishing (AW NRM and deeper waters seaward) Reduce the quota for Southern Rock Lobster in the Northern Zone, in light of the apparent long-term decline in the stock. High Fishing (AW NRM and deeper waters seaward of AW NRM) Sharks and other elasmobranch species listed internationally as vulnerable or endangered, should be fully protected from commercial and recreational fishing in the Great Australian Bight, including waters of the AW NRM. This includes school shark, thresher shark, black whaler, shortfin mako, smooth hammerhead, porbeagle, Melbourne skate, and coastal stingaree. High Fishing (AW NRM and deeper waters seaward of AW NRM) A seasonal restriction on fishing gummy sharks should be considered during the pupping season. This should apply to both commercial and recreational fishing. High 96 Table 8 (cont.): Summary of issues / threat categories, area of concern (within AW NRM and/or seaward of AW NRM region), proposed actions, and priority ranking for actions. Issue / Threat Category (and Area of Concern) Proposed Action Fishing (AW NRM and deeper waters seaward of AW NRM) Increase the level and quality of quantitative data on bycatch rates of Sea Lions in the South Australian Rock Lobster fishery. Anecdotal evidence and entanglement data suggest there has been significant underreporting of seal and sea lion interactions in this fishery (SARDI data, 2006). Priority of Action In both Western Australia and South Australia, research is currently being undertaken to determine suitable methods to modify rock lobster pots to eliminate capture of juvenile Australian sea lions, to reduce humaninduced mortality in the population (Campbell, 2005; Goldsworthy, 2008), Measures such as seal-exclusion devices (a metal spike placed vertically in the neck of the pot), and attaching bait containers to the side of the pot entrance, amongst other measures, should be adopted in all parts of the South Australian rock lobster fishery. Further information on reducing interaction of sea lions within the rock lobster industry is provided in Goldsworthy (2008). Fishing (deeper waters seaward of AW NRM) Reducing the bycatch of attached benthos such as sponges and hard corals will reduce habitat degradation and this may help to ensure the healthy status of fish stocks in the Great Australian Bight. This may be achieved in the closing areas of ―virgin‖ (untrawled) ground to trawling and by prohibiting exploratory fishing under certain depths (<100m). Consideration may also be given to gear types or modifications that are less damaging to benthic environments (Brown and Knuckey, 2002). Periodically closing trawled areas (particularly those for which sponge and/or coral bycatch is large) to permit recovery should also be considered. Benthic habitats dominated by dominated by large sponges and mixed epifauna, and those that are dominated by bryozoan communities (e.g. at the shelf break) should be closed to trawling. High High Fishing (deeper waters seaward of AW NRM) Investigate means of reducing bycatch of non-target species, in terms of timing of shots (moon phase etc) and location of shots, as well as further gear modifications. Also, avoid fishing areas where catches of species of conservation concern (e.g. stingarees, sharks, dogfishes) are highest. High Fishing (deeper waters seaward of AW NRM) Reduce risk of entanglement of cetaceans (and sea lions – see separate heading) in trawl, gillnet and longline fisheries, by gear modification, seasonal or permanent closures in areas where cetaceans are known to aggregate, and acoustic alarms on vessels during whale migration season. High 97 Table 8 (cont.): Summary of issues / threat categories, area of concern (within AW NRM and/or seaward of AW NRM region), proposed actions, and priority ranking for actions. Issue / Threat Category (and Area of Concern) Proposed Action Fishing (deeper waters seaward of AW NRM) Investigate the extent of the deliberate harming of both sharks and cetaceans by fishers operating seasonally in the GAB, and where possible, ensure that on-board observers are present to prevent such practices. High Threats to Sea Lions (AW NRM) Prohibit entry by land or sea to any of the Bunda Cliff colonies, with the exception of non-invasive research by permit. High Threats to Sea Lions (deeper waters seaward of AW NRM) Based on the outcome of current research which is investigating the main foraging areas of sea lions in south Australia, restrict trawl fishing and gillnetting in the main foraging areas of sea lions, including area in the Great Australian Bight. This may be particularly important during breeding times when the young are left ashore to await a parent‘s return from feeding expedition. High Increase the level and quality of quantitative data on bycatch rates of sea lions in the shark gillnet sector of the SESSF. Anecdotal evidence and entanglement data suggest there has been significant underreporting of seal and sea lion interactions in this fishery (SARDI data, 2006). Recreation and Tourism (AW NRM) Reduce the number of vehicle tracks and foot access tracks through the dunes and other coastal habitats in the AW NRM region. Reduce or restrict the use of vehicles on sandy beaches which contain beach wrack, particularly during the hooded plover breeding season. Beaches where this vulnerable bird species is known to nest i.e. where nesting is confirmed, should be closed to vehicles. Divert cliff-top access tracks and roads away from osprey and white-bellied sea eagle nests. Dennis (2008) discussed this requirement in more detail (particularly for the Yalata Indigenous Protected Area between Head of the Bight whale viewing area and the boundary of the Nullarbor National Park), and his suggestions are fully supported here. Restrict access by visitors / tourists to Head of Bight locations where active osprey and white-bellied sea eagle nests occur. Reduce the opportunity for off-road vehicles to make informal tracks along the cliffs to viewing spots. Prohibit the dumping of rubbish by visitors at accessible cliff and beach sites. Ensure that visitors to the Head of the Bight to view whales and sea lions adhere to the standards from the policy associated with recreational and commercial tourism interactions with marine mammals in South Australia (see National Seal Strategy Group and Stewardson, 2005). 98 Priority of Action High High High High (for all recommendations) Table 8 (cont.): Summary of issues / threat categories, area of concern (within AW NRM and/or seaward of AW NRM region), proposed actions, and priority ranking for actions. Issue / Threat Category (and Area of Concern) Marine Debris (AW NRM and deeper waters seaward of AW NRM) Proposed Action Priority of Action All NRM Boards and government agencies with marine responsibilities should be aware of the national threat abatement plan for the impacts of marine debris on vertebrate marine life (DEWHA, 2009a). Although South Australia is only briefly considered in the plan, many of the proposed actions are relevant to marine debris reduction in all States and Commonwealth waters. Adherence to the plan will also ensure that operators in Commonwealth and southern Australian State waters remain compliant with the international MARPOL (Annex V) regulations. High (for all recommendations) Where possible, expand South Australian investigation of the spatial and temporal extent of mortality and injury of marine biota (particularly marine mammals) due to fishing industry discards, to include the fished area of the GAB. Prohibit the use of plastic bait box straps in all Australian fisheries, including the shark and long-line tuna fisheries. Marine Pests (AW NRM) Although there are no harbours in the AW NRM region, and no shipping activity (see above), even small recreational and commercial boats can carry marine pests. Therefore, vessels entering waters at the Head of the Bight should be cautious in this regard, and take steps to ensure that ballast water is not discharged. Hulls should be cleaned regularly, with no discharge of waste washing waters to sea. High (for all recommendations) Recreational boats and associated equipment (e.g. anchors) from Gulf St Vincent should not be permitted to enter the waters of AW NRM, to reduce the risk of the invasive Caulerpa taxifolia being transferred to the region. Researchers working in the GAB Marine Park have been asked to report any sightings of marine pests (Director of National Parks, 2005). Research (AW NRM) Researchers undertaking work within sea lion colonies across the GAB should minimise disturbance, as recommended by the Director of National Parks (2005). High Research (AW NRM, and deeper waters seaward of AW NRM) Benthic researchers should minimise the use of benthic sleds, dredges and cores, because these methods can damage benthic biota, and degrade habitat quality. High Land-based and Coastal Discharges (AW NRM) Reduce the proliferation of tracks in some parts of the cliff top and cliff face within the AW NRM region, to reduce gully erosion. High Reduce the number of tracks through dunes, to try to mitigate dune destabilisation and erosion (see also Caton et al., 2007) 99 Table 8 (cont.): Summary of issues / threat categories, area of concern (within AW NRM and/or seaward of AW NRM region), proposed actions, and priority ranking for actions. Issue / Threat Category (and Area of Concern) Proposed Action Priority of Action Mining, Minerals, Oil and Gas Exploration (deeper waters seaward of AW NRM) Prohibit seismic exploration, drilling, and all other forms of exploration for oil, gas, mining and minerals in the Great Australian Bight Marine Park, and in any future extension of that park (e.g. see section 3.3 below, on proposed GAB Extension, as part of the Commonwealth Marine Reserves program). High Climate Change The effects of climate change in any region cannot easily be detected or monitored without knowledge of the distribution, abundance and species composition of marine flora and fauna. Therefore, more research is required to document these features as a baseline against which future changes can be measured. There are many knowledge gaps in the AW NRM (some in the process of being filled), with some examples including baseline data on ranges of marine macroalgae, seagrasses, benthic and demersal fishes, attached and mobile invertebrates, temperature tolerances, and demography. High Climate Change Following baseline data collection, monitoring at specific time intervals for all of the above groups is required to assess climate change impacts. This would include programs to monitor species identified as ‗indicators‘ of climate change (Hobday et al., 2006). Long term monitoring programs are crucial to an understanding of impacts, yet in Australian marine research, there is still a significant underinvestment in long-term monitoring (Hobday et al., 2006). High Climate Change Ensure that the South Australian component of IMOS, the Integrated Marine Observing System, continues to include sites in the GAB for long term monitoring, and that such monitoring include sea surface temperature, salinity, currents, water chemistry, seabed habitats, phytoplankton, zooplankton and acoustic monitoring of marine animal movements. Medium Expand the IMOS system to include inshore waters. Climate Change Develop models to assess how main habitat types, such as rocky shore communities, subtidal reefs (dominated by kelp and other canopy-forming macroalgae) and invertebratedominated communities may change in the future. In Australia, greater emphasis on data analysis and modelling is required in future to better predict specific climate change impacts , rather than just warehousing and data-basing large amounts of unused biophysical inventories (Rosenzweig et al., 2008). 100 Medium Table 8 (cont.): Summary of issues / threat categories, area of concern (within AW NRM and/or seaward of AW NRM region), proposed actions, and priority ranking for actions. Issue / Threat Category (and Area of Concern) Proposed Action Climate Change Although climate change effects cannot be mitigated over the short and medium term, providing opportunity for climate changestressed organisms to maintain adaptive responses, includes means of controlling other interactive impacts by the following (from Babcock et al., in Poloczanska et al., 2008; Booth et al., 2009 and Wernberg et al., 2009): Priority of Action High (i) reducing fishing levels (and ensuring that stocks are not driven down to low levels); (ii) formally protecting habitat and strongly site-associated species (through sanctuary zones, in Marine Protected Areas). (iii) utilising Marine Protected Areas to assist in providing fish populations with some level of resistance to climate change, by allowing populations to exist with reduced fishing pressure; (iv) restocking programs for key species, if careful attention is paid to genetic structure and indirect effects on local marine communities. (v) adaptive fisheries management, such as offering incentives to fishers to target species that are less affected by climate change; concomitant incentives for consumers to switch to such species as preferred food choices; and considerably reducing catch levels of commercial species that are most affected by climate change; and (vi) ensuring that coastal discharges and other agents of increased nutrients and other pollutants in marine systems are controlled, given the evidence that climate change impacts in marine ecosystems can be exacerbated by other (humaninduced) stressors. (vii) reducing the opportunities for physical, chemical and biological disturbance to habitats (e.g. by prohibiting coastal dredging and developments, and other disruptive works). Other The knowledge of indigenous custodians of the area should be incorporated into conservation and management strategies, and these people should be active participants in such strategies, and in long-term regional management plans. 101 High Table 8 (cont.): Summary of issues / threat categories, area of concern (within AW NRM and/or seaward of AW NRM region), proposed actions, and priority ranking for actions. Issue / Threat Category (and Area of Concern) Proposed Action Climate Change Wernberg et al. (2009) provided suggestions for more experimental work (including integrated field and laboratory manipulations) to (i) increase knowledge of the current changes to temperate reef assemblages, and (ii) better predict future changes. Medium Climate Change Employ ecosystem models to distinguish climate change impacts from other impacts (e.g. fisheries impacts). Richardson and Okey, in Hobday et al. (2006) and Okey et al., in Poloczanska et al. (2008), provided a simple summary of types of model approaches that may be useful for describing and predicting climate change impacts. Medium Climate Change The likely interaction between climate change variables and fishing practices needs investigating (Booth et al., 2009). Medium Climate Change Electronic tagging of indicator species, such as tuna and sharks, to develop habitat preference models, and to assist in determining range shifts over time (e.g. tunas are predicted to move southwards with oceanic warming) (Hobday et al., 2006). Medium 102 Priority of Action Part 4 CONSERVATION AND MANAGEMENT IN THE AW NRM REGION – MARINE COMPONENT 4.1 Notes on Current Protection and Management The waters of the AW NRM region are in a unique position in terms of protected area legislation, because they are included in the State waters section of the Great Australian Bight Marine Park (as discussed below), and the coastal lands backing those waters are part of the Nullarbor National Park and the Yalata Indigenous Protected Area (Figure 12). More recently (2010), the Nullarbor Plain was declared a Wilderness Protection Area, under the Wilderness Protection Act 1992. Figure 12: Marine and terrestrial areas protected under legislation in and adjacent to the AW NRM region (Figure from Australian Government DEH, 2005). The Nullarbor National Park (declared 1979) and Regional Reserve (proclaimed 1989), aim to protect the world's largest semi-arid karst (cave) landscapes and associated wildlife, including the southern hairy-nosed wombat. However, these parks do not have a high level of area protection against activities such as mining, compared with Wilderness Protection Areas in South Australia. (see below). The recent declaration of the Nullarbor Plain as a Wilderness Protection Area (WPA) under the Wilderness Protection Act 1992, follows a long-term campaign and nomination by the Wilderness Society. The nomination acknowledges the cultural significance of the region to the Mirning people and their interests in its protection and future management. It also recognises the international and national importance of the area for its natural and cultural heritage values, as the largest semi-arid karst cave system in the world, covering an area of more than 200,000km. The spectacular Bunda Cliffs and the Great Australian Bight bordering the area to the south, are included in the WPA, which will ostensibly increase the level of protection of Australian Sea Lion colonies. Also, no mining exploration or mining activity will be permitted in the WPA. 103 In the coastal area of the Head of the Bight, a permit system is used to access the Yalata Indigenous Protected Area. In that area, the impact of the steadily increasing number of people visiting Yalata's beaches has reportedly become more significant in recent years (Yalata Land Management, 2003), and one strategy to help control impacts of recreational activities is to control numbers of visitors through the permit system. The funds collected from permits aid the land management activities in the area, as well as providing rangers to ensure that the coastal environment is protected, and the health and safety of visitors is overseen (Yalata Land Management, 2003). In the coastal dune areas, vehicles must use designated tracks only (as written on the permit); access through dunes is via marked roads only; camping must be in designated areas only (as specified on the permit); pets must not roam free; all rubbish from campsites must be removed; firearms are not permitted to be used on IPA lands, and permit holders must not interfere with Aboriginal cultural heritage, flora and fauna (Yalata Land Management, 2003). Adjacent to the coast, the waters of the AW NRM region fall within the State water component of the Great Australian Bight Marine Park. The Park is made up of adjoining South Australian and Commonwealth protected areas (Figure 13). The State Marine Park, in the State (coastal) waters of the Bight, combines a whale sanctuary established under the former South Australian Fisheries Act 1982 (now the Fisheries Management Act 2007) and marine national park established under the National Parks and Wildlife Act 1972. The sanctuary and marine national park in S.A. waters were declared during the mid 1990s. The adjoining Great Australian Bight Marine Park (Commonwealth Waters) is a Commonwealth reserve managed under the Environment Protection and Biodiversity Conservation Act 1999 (and established the former Commonwealth National Parks and Wildlife Conservation Act 1975). The relevant Acts contain penalties for offences such as entering prohibited areas or conducting prohibited activities within the Park. The reserved area in State waters includes a Sanctuary Zone and a Conservation Zone (Figure 13). The Commonwealth waters component (seaward of AW NRM region) includes a Marine Mammal Protection Zone adjacent to State waters, and a Benthic Protection Zone strip, consisting of a 20 NM-wide strip orientated north to south and extending from three nautical miles from the coast to the edge of Australia‘s EEZ, 200 nautical miles offshore (Figure 13). Collectively, the zones are ostensibly designed to protect the main conservation values of the Park, which are considered by government to be (i) habitat for the southern right whale (Eubalaena australis). Protection of the southern right whales and their habitat in the Great Australian Bight Marine Park is considered important for the global conservation of this species; (ii) habitat for the Australian sea lion (Neophoca cinerea); (iii) habitat for other species of conservation significance, and (iv) a transect representative of the seabed on the continental shelf and slope of the GAB (Australian Government DEH, 2005). 104 Figure 13: Great Australian Bight Marine Park, showing State and Commonwealth components. Figure from Australian Government DEH (2005). The park is zoned, with different rules in each zone, in order to specific species. These zones are shown in Table 9 below. Beach based recreational fishing can only take place in the sanctuary zone of the Park and any boating activity must conform to the zoning conditions. The normal State regulations under the Fisheries Management Act 2007 still apply in relation to legal lengths, catch limits, permitted gear, and any other conditions. The zoning regulations in the GAB Marine Park are discussed in more detail in Department for Environment and Heritage‘s (1998) Great Australian Bight Marine Park Management Plan, Part A. 105 Table 9: Zones of the Great Australian Bight Marine Park in South Australian- and Commonwealth-managed waters, and permitted activities (from Department for Environment and Heritage, 2009). Zone Purpose Rules Protecting Southern Right Whales, Australian Sea Lions & representative habitat for Southern Right Whales No boating or resource extractive activities* are allowed (except for recreational line fishing from the beach). Conservation Zone (State) Protecting Southern Right Whales & Australian Sea Lions Annual six month closure from May to October every year to all boating activity (covers the period the whales are present). All boating and commercial fishing activity allowed at other times. Marine Mammal Protection Zone (Commonwealth) Protecting Southern Right Whales & Australian Sea Lions Annual six month closure from May to October every year to all boating activity (covers the period the whales are present.) Sanctuary Zone (State) * (includes mining, bio-prospecting, commercial fishing, recreational fishing from boats, mining activities etc). All boating and commercial fishing activity allowed at other times, except trawling. Benthic Protection Zone (Commonwealth) Protecting the sea floor habitat structure and species that live on the sea floor (because of the unique nature of the assemblages present, with many southern Australian endemic species) No demersal (sea floor) trawling allowed (as this activity alters the seafloor structure). ―More stringent approval process for petroleum exploration‖. In the GAB, there are several closures in place (Figure 14) which restrict commercial fishing, and in the shallow shelf waters of the AW NRM region, the GAB Marine Park is relevant in this regard. The shallow waters of the GAB Marine Park Mammal Protection Zone are not in the area of the Commonwealth trawl fisheries. A research and management strategy for the Commonwelath and State waters of the GAB Marine Park was being written in June 2010 (S. Kumar, DEH, pers. comm., 2010). 106 Figure 14: Areas closed to fishing, or which fishing is seasonally restricted, in the Great Australian Bight. Map from GABIA, reproduced from AFMA, 2008C). In deeper shelf waters, in the GAB trawl sector of the SESSF, quota management of the main target species was introduced in 2006, comprising equal allocation of quota between ten Statutory Fishing Rights licence holders. Selectivity of the gear used in the GABTF is governed by a minimum mesh size of 90mm, but due to an industry code of conduct, mostly 100mm minimum mesh sizes are used in the fishery. Over the past few years the GABTF has invested in developing and trialling fishing gear to improve selectivity of target species and to reduce bycatch. The GAB Industry Association agreed that all vessels in the fishery must use T90 extensions in all demersal nets, which result in a reduction in the catch of small unwanted species. Testing of further developments in net design and gear configurations is currently being conducted under scientific conditions to qualify and quantify the bycatch reductions and selectivity (AFMA, 2008c). Other developments include assessing the recording of bycatch in logbooks every 3-6 months; and development and implement of individual vessel management plans (VMP), to minimise seabird interactions (e.g. offal management), and record procedures for reporting on catch and wildlife interactions. As part of that aim, every vessel must design and carry an up-to-date VMP, to be recorded by observers (AFMA, 2008c). Spatial closures to fishing have been implemented in areas of the GAB where pregnant female school sharks are known to aggregate (AFMA, 2008b). Another current management strategy in the commercial shark fishery in southern Australia is to protect a portion of the large breeding females of gummy shark. 107 4.2 Notes on Proposed Protection and Management During the mid to late 2000s, the Department of Environment and Heritage in South Australia developed a proposal to declare a multiple use marine park in Head of the Bight region, between the Western Australian border and the Tchalingaby sandhills. The Far West Coast Marine Park includes the State waters of the Great Australian Bight Marine Park (DEH, 2009). The Far West Marine Park comprises the area bounded by a line commencing on the coastline at median high water at a point 132°0´4.97E´´, 31°52´38.75S´´ then running progressively:  southerly along the geodesic to its intersection with the seaward limit of the coastal waters of the State at a point 132°0‘4.97‖E, 31°56‘16.08‖S;  generally westerly along the seaward limit of the coastal waters of the State to a point 129°0‘5.08‖E´, 31°44‘29.80‖S;  northerly along the geodesic to its intersection with the coastline at median high water at a point 129°0‘5.08‖E, 31°41‘12.91‖S (at the western edge of Nullarbor National Park); and  generally easterly along the coastline at median high water (inclusive of all bays, lagoons and headlands) to the point of commencement (DEH, 2009). Zoning (including high protection sanctuary zones) for the multiple use marine park had not been determined at the time of writing this report. In deeper, Commonwealth-managed waters, a marine bioregional planning program started in 2006, for a broad area known as the South-West Marine region, spanning from Kangaroo Island in South Australia through to Shark Bay in Western Australia. The program will result in the development of marine bioregional plans under national environment law for all Australian waters. It will also establish networks of Commonwealth Marine Reserves which will contribute to the National Representative System of Marine Protected Areas (NRSMPA). All governments in Australia have a shared a national and international commitment to establish a national representative system by the year 2012 (DEWHA, 2009b). One of the 7 areas for further assessment is the Head of the Great Australian Bight, including a proposed extension of the current GAB Marine Park, in Commonwealth-managed adjacent to the AW NRM marine boundary (see Figure 15 below). 108 Figure 15: One of the 7 major areas for further assessment in the South West Marine Region (from DEWHA, 2009b) The ―Great Australian Bight (Extension)‖ area for further assessment is an extension of both the inner shelf and benthic protection components of the existing Great Australian Bight Marine Park. The proposed reserve area recognises (i) the notion of the Head of the Bight being a ―hotspot‖ of biodiversity and productivity; (ii) the importance of this area for invertebrate communities of the inner shelf, particularly the diversity of sponges, ascidians and bryozoans; (iii) the national importance as a calving area for Southern Right Whales; and (iv) the important of the area for other protected species, such as the Australian Sea Lion and the White Shark. The proposed conservation objectives for any reserve(s) established within the ―GAB (extension) Area for Further Assessment‖ area are as follows (from DEWHA, 2009b): (i) protect ecologically adequate examples of benthic invertebrate habitat communities of the Great Australian Bight, specifically within the Eucla meso-scale bioregion and the Southern Province; (ii) to protect ecologically adequate examples of the pelagic environments associated with the Great Australian Bight Shelf Transition and the Southern Province; (iii) to preserve in their unmodified and/or natural conditions, areas of the Great Australian Bight Shelf Transition and the Southern Province; (iv) to contribute to the recovery and long-term protection of Southern Right Whales, and (v) to contribute to the recovery and long-term protection of the Australian Sea Lion, including the species‘ genetic diversity (DEWHA, 2009b). 109 5. Acknowledgements Thanks to Commonwealth government‘s Caring for our Country for providing the funding for this report to be written. The Alinytjara Wilurara Natural Resources Management Board engaged J.L. Baker, Marine Ecologist write the marine threats review. The author would like to acknowledge and thank A. Loisier (AW NRM Board) for providing editorial advice in the development of this document, and for providing the table on Hooded Plover survey numbers. Thanks also to A. Wood and A. Loisier at AW NRM Board for providing several of the maps. 6. Bibliography ABRS (Australian Biological Resources Study) (2009) Australian Faunal Directory. Department of the Environment, Water, Heritage and the Arts (DEWHA), Canberra. ABS (Australian Bureau of Statistics) (2007) Year Book Australia, 2005. Transport Infrastructure. http://www.abs.gov.au/AUSSTATS/abs@.nsf/Previousproducts/FA7086AC621CD2DFCA256F7200832F C2?opendocument AFFA (2002) Fishery Status Reports 2000-2001. Australian Government Department of Agriculture, Fisheries and Forestry, Canberra. AFMA (2002) Proposed Area Closures in the Southern Shark and South East Non-Trawl Fisheries. Discussion Paper, July 2002. Australian Fisheries Management Authority, Canberra. AFMA (2003) SESSF Strategic Assessment Report. Australian Fisheries Management Authority, Canberra. AFMA (2003) Southern and Eastern Scalefish and Shark Fishery Management Plan 2003. Australian Fisheries Management Authority, Canberra. AFMA (2006a) Southern and Eastern Scalefish and Shark Fishery - Great Australian Bight Trawl Management Advisory Committee (GABMAC). Great Australian Bight Trawl Fishery ISMP Observer Report (Spring 2005/Summer, Autumn 2006). Attachment to Minutes of Meeting held July, 2006. Australian Fisheries Management Authority, Canberra. AFMA (2006b) Southern and Eastern Scalefish and Shark Fishery - Great Australian Bight Trawl Management Advisory Committee (GABMAC). Minutes of meeting, September, 2006. Australian Fisheries Management Authority, Canberra. AFMA (2008a) Ecological Risk Management: Level 2 ERA Residual Risk Assessments for Commonwealth Fisheries. http://www.afma.gov.au/environment/eco_based/eras/res_assessments.htm AFMA (2008b) Ecological Risk Management. Report for the Great Australian Bight Trawl sub-fishery of The Southern and Eastern Scalefish and Shark Fishery. December, 2008. Australian Fisheries Management Authority, Canberra. AFMA (2008c) Great Australian Bight Trawl Fishery: Bycatch and Discarding Workplan November 2008. Australian Fisheries Management Authority, Canberra. 110 Allen, S. and Bejder, L (2003) Southern right whale (Eubalaena australis) sightings on the Australian coast and the increasing potential for entanglement. Pacific Conservation Biology 9 (3). Anonymous (2000) Australia’s Great Road Journey: the Nullarbor. Brochure produced by Eyre Peninsula Tourism Association, Eyre Highway Operators Association, South Australian Tourism Commission, Esperance Region Tourism Association and Goldfields Tourism Association. ANRA (Australian Natural Resources Atlas) (2009) Rangelands – Overview. http://www.anra.gov.au/topics/rangelands/overview/sa/ibra-nll.html. Department of the Environment, Water, Heritage and the Arts. APPEA (undated) A compilation of recent research into the marine environment. Report by the Australian Petroleum Production and Exploration Association, Canberra. Atlas of South Australia (2005) http://www.atlas.sa.gov.au Australian Government Department of the Environment and Heritage (2004) Assessment of the ecological sustainability of management arrangements for the South Australian Pilchard Fishery. Australian Government DEH, Canberra. Australian Government Department of the Environment and Heritage (DEH) (2005) Great Australian Bight Marine Park (Commonwealth Waters and State Waters) – A Description of Values and Uses. Australian Government Department of the Environment and Heritage, Canberra. Australian Government Department of the Environment, Water, Heritage and the Arts (2009) Species Profile and Threats Database. http://www.environment.gov.au/cgi-bin/sprat/public/sprat.pl Australian Heritage Commission (AHC) (undated). Website: http://www.ahc.gov.au/heritage/register/easydatabase/database.html. Australian Museum (2003) Spotted Wobbegong Orectolobus maculatus. Web page compiled by M. McGrouther. http://www.amonline.net.au/fishes/students/focus/gwobbe.htm Australian Museum (2003) Ornate Wobbegong Orectolobus ornatus. Web page compiled by M. McGrouther. http://www.amonline.net.au/fishes/fishfacts/fish/oornatus.htm Australian Museum (2009a) Short-finned Conger-eel, Conger wilsoni (Bloch & Schneider, 1801) http://australianmuseum.net.au/Short-finned-Conger-eel-Conger-wilsoni-Bloch-Schneider-1801 Ayling, T. and Cox, G. (1982) Collins Guide to the Sea Fishes of New Zealand. Collins, Auckland, New Zealand. 343p. Bail, P. and Limpus, A. (1997) (description of Paramoria johnclarki) Apex 12(4): 109-115. Baker, J.L. (2004) Towards a System of Ecologically Representative Marine Protected Areas in South Australian Marine Bioregions - Technical Report. Report for Coast and Marine Conservation Branch, Department for Environment and Heritage (DEH), South Australia. 1350p. (CD version available from Coast and Marine Conservation Branch, DEH, South Australia. Ph.(08) 8124 4900). http://www.environment.sa.gov.au/marineparks/pdfs/part_2.pdf http://www.environment.sa.gov.au/marineparks/pdfs/part_3.pdf 111 Baker, J.L. (2007a) Inshore fish. Chapter in: McClatchie, S. et al. (Eds) Review of Ecological Information and Knowledge of Australia‘s South West Marine Region (SWMR). Report by SARDI Aquatic Sciences and University of Western Australia, for National Oceans Office, Canberra. http://www.environment.gov.au/coasts/mbp/publications/south-west/pubs/sw-ecosystems-part2.pdf Baker J.L. (2007b) Syngnathids. Chapter in: McClatchie, S. et al. (Eds) Review of Ecological Information and Knowledge of Australia‘s South West Marine Region (SWMR). Report by SARDI Aquatic Sciences and University of Western Australia, for National Oceans Office, Canberra. http://www.environment.gov.au/coasts/mbp/publications/south-west/pubs/sw-ecosystems-part2.pdf Baker, J.L. (2009-2010) Status of Marine Species of Conservation Concern in South Australia: Technical Report – Bony and Cartilaginous Fish. Report and CD prepared for the South Australian Working Group for Marine Species of Conservation Concern. J. Baker (consultant); Science and Conservation Division, and Coast and Marine Conservation branches of S.A. Department for Environment and Heritage (DEH); Marine and Coastal Community Network of S.A. (MCCN); Reef Watch, South Australia, and Threatened Species Network (TSN). Baker, J.L., Rodda, K.R. and Shepherd, S.A. (2008) Sharks and rays of Gulf St Vincent. In: Shepherd, S.A., Bryars, S., Kirkegaard, I., Harbison, P. and Jennings, J. (Eds) The Natural History of Gulf St Vincent. Royal Society of South Australia. Bannister J.L. (1979) Southern right whale aerial survey and photoidentification, southern Australia. Annual reports 1979-2005. Australian Government Department of Environment and Heritage, Canberra. (Unpublished report). Barrow, R.A. and Capon, R.J. (1994) Carduusynes (-E): Acetylenic Acids from a Great Australian Bight marine sponge Phakellia carduus. Australian Journal of Chemistry. 47(10): 1901-1918. Bax, N. and Williams, A. (2001) Delineating fish-habitat associations for spatially based management: an example from the south-eastern Australian continental shelf. Marine and Freshwater Research 52: 513 – 536. Beechey, D. (2009) The Seashells of New South Wales. Based on the collections of the Australian Museum and of the author (Des Beechey, Senior Fellow, Australian Museum). Release 14, October 2009. Bester, C. (2003) Biological Profiles: Tope Shark. Ichthyology Section, Florida Museum of Natural History: http://www.flmnh.ufl.edu/fish/Gallery/Descript/TopeShark/TopeShark.html Beu, A.G. (2004) Marine Mollusca of oxygen isotope stages of the last 2 million years in New Zealand. Part 1: Revised generic positions and recognition of warm-water and cool-water migrants. Journal of the Royal Society of New Zealand 34(2): 111-265. Bock, P.E. (1982) Bryozoans. Chapter 9 in: Shepherd, S. and (the late) Thomas, I. (Eds) Marine Invertebrates of Southern Australia, Part 1. Handbook of South Australian Flora and Fauna, Government Printer, Adelaide, South Australia. Bone, Y. (1997) Bryozoa: living biota to sedimentary fragments to limestones: How do oceanographic parameters of the Great Australian Bight affect this progression? Oceanography of the South East Indian Ocean and Great Australian Bight Conference, Oceanography Workshop, Flinders University, South Australia, Sept., 1997, Technical Report No. 14, pp. 16-17. Bone, Y. (1998) Cool-water carbonate sedimentation, Bonney and Lacepede Shelves and Eastern Great Australian Bight. ORV Franklin cruise summary. University of Adelaide report, for CSIRO, Hobart. 112 Bone, Y. and James, N.P. (1998a) Biota and sediments on the Great Australian Bight: Their interdependence, deposition and potential for accumulation. The South East Indian Ocean and Great Australian Bight USA/Australia Bilateral Workshop, Marine Research Centre, Pt. Lincoln, Sept., 1998, 1st Edition, pp. 4-5. (invited address) Bone, Y. and James, N.P. (1998b) Oceanography and hydro-geochemistry of the Great Australian Bight: preliminary results. Invited lecture given at Great Australian Bight Workshop, Environment Australia, Kangaroo Island, Oct. 1998. Bone, Y. and James, N.P. (2002) Bryozoans from Pleistocene high-energy, cool-water, continental-slope mounds, Great Australian Bight, southern Australia. Abstracts - Geological Society of Australia No. 67. p. 352. Booth, D., Edgar, G., Figueira, W., Jenkins, G., Kingsford, M., Lenanton, R. and Thresher, R. (2009) Temperate coastal and demersal fish and climate change. In: In: Poloczanska, E.S., Hobday, A.J. and Richardson, A.J. (Eds) A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009. NCCARF Publication 05/09, ISBN 978-1-921609-03-9. Bornatowski, H., Costa, L., Robert, M.C. and Pina, J.V. (2007) Hábitos alimentares de tubarões-martelo jovens, Sphyrna zygaena (Carcharhiniformes: Sphyrnidae), no litoral sul do Brasil. Biota Neotrop. 7(1): 213-216. Bradbury, J.H. and Eberhard, S. (2000) A new stygobiont melitid amphipod from the Nullarbor Plain. Records of the Western Australian Museum 20: 39-50. Branden, K.L. and Shepherd, S.A. (1984) A survey of stocks of the abalone Haliotis roei Gray in the north-eastern Great Australian Bight. Fisheries Research Paper No. 11. Department of Fisheries (South Australia). 10p. Bratcher, T. and Cernohorsky, W. (1987). Living Terebras of the World. A Monograph of the Recent Terebridae of the World. American Malacologists, Melbourne, Florida. 240p. Britz, P.J., Sauer, W.H.H., Mather, D., Oellerman, L., Cowley, P.D., Ter Morshuizen, L. and Bacela, N. (2001) Baseline study of the utilisation of living marine resources in the Eastern Cape Province. Report for the Department of Economic Affairs, Environment and Tourism, Eastern Cape Province. Department Of Ichthyology & Fisheries Science, Rhodes University, South Africa. Brown, L. and Knuckey, I. (2002) ISMP – Monitoring of the Great Australian Bight Trawl Fishery during 2001/02. Report to Fisheries Research and Development Corporation, Project No. R01/0817. Department of Natural Resources and Environment, Victoria. Brown, L., Bridge, N. and Walker, T. (2000) Summary of tag releases and recaptures in the Southern Shark Fishery. Marine and Freshwater Resources Institute Report No.16. 61p. Bruce, B.D., Malcolm, H. and Stevens, J.D. (2001) A Review of the Biology and Status of White Sharks in Australian Waters. CSIRO Marine Research, Hobart, Tasmania. Bruce, B.D., Stevens, J.D. and Malcolm, H. (2002) Broad-scale movements of white sharks (Carcharodon carcharias) in Australian waters. Presented at the Australian Society for fish Biology Annual Conference, 2002. Bruce, B.D., Stevens, J.D. and Malcolm, H. (2006) Movements and swimming behaviour of white sharks (Carcharodon carcharias) in Australian waters. Marine Biology 150: 161-172. 113 Bruce, B.D., Bradford, R., Daley, R.K., Green, M. and Phillips, K. (2002) Targeted review of biological and ecological information from fisheries research in the South East Marine Region. Final Report. CSIRO Marine Research report for the National Oceans Office, Canberra. Bureau of Meteorology (2010) Climate variability and El Niño. http://www.bom.gov.au/lam/climate/levelthree/analclim/elnino.htm Burnell, S.R. (1999) The population biology of southern right whales in southern Australian waters. PhD thesis, University of Sydney, New South Wales. Burnell, S. (2001) Aspects of the reproductive biology, movements and site fidelity of right whales off Australia. Journal of Cetacean Research and Management (Special Issue) 2: 89-102. Burnell, S. and Pirzl, R. (2006) Head of Bight Right Whales: Research Overview. Presentation at the DEH Cetacean Research Priorities Conference, Adelaide SA, February 2006. Deakin University. Campbell, R. (2003) Demographic and population genetic structure of the Australian sea lion, Neophoca cinerea. PhD Thesis. University of Western Australia, Perth, WA. Campbell, R. (2005) Historical distribution and abundance of the Australian sea lion (Neophoca cinerea) on the west coast of Western Australia, Fisheries Research Report No. 148, Department of Fisheries, Western Australia, 42p. Cane, S. (1989) UNDIRI: Aboriginal Association with the Nullarbor Plain. A Report to the National Parks and Wildlife Service, South Australia. Capon, R.J., Miller, M. and Rooney, F. (2000) Clathrins A-C: metabolites from a southern Australian marine sponge, Clathria. Journal of Natural Products 63 (6): 821-824. Capon, R.J., Groves, D.R., Urban S. and Watson R.G. (1993) Spongiaquinone revisited: structural and stereochemical studies on marine sesquiterpene/quinones from a southern Australian marine sponge, Spongia sp. Australian Journal of Chemistry 46: 1245-1253. Capon, R.J., Peng, C. and Dooms, C. (2008) Trachycladindoles A–G: cytotoxic heterocycles from an Australian marine sponge, Trachycladus laevispirulifer. Org. Biomol. Chem. 6: 2765 – 2771. Cappo, M. (1992) Bronze Whaler sharks in South Australia. Safish Magazine July–September 1992: 10– 13. Caputi, N., de Lestang, S., Feng, M. and Pearce, A. (2009) Seasonal variation in the long-term warming trend in water temperature off the Western Australian coast. Marine Freshwater Research 60: 129-139. Carpenter, G. (2005) Birds of the Merdayerrah region, Nullarbor Plain, SA. Unpublished report., South Australian Museum. Carraro, R. and Gladstone, W. (2006) Habitat preferences and site fidelity of the ornate wobbegong shark (Orectolobus ornatus) on rocky reefs of New South Wales. Pacific Science 60: 207-223. Caton, A.E. (Ed.) (2003) Fishery Status Reports 2002–2003: Assessments of the Status of Fish Stocks Managed by the Australian Government. Bureau of Rural Sciences, Canberra. Caton, B., Detmar, S., Fotheringham, D., Haby, N., Royal, M., and Sandercock, R. (2007) Far West Coastal Action Plan and Conservation Priority Study. Prepared for Alinytjara Wilurara NRM Board and Department for Environment and Heritage. 114 Cavanagh, R.D., Kyne, P.M., Fowler, S.L., Musick, J.A. and Bennett, M.B. (2003) The Conservation Status of Australasian Chondrichthyans: IUCN Shark Specialist Group Australia and Oceania Regional Red List Workshop Report, Queensland, 7-9 March 2003. University of Queensland, Brisbane, Queensland. Ceccarelli, D. M. (2009) Impacts of plastic debris on Australian marine wildlife. Report by C&R Consulting for the Department of the Environment, Water, Heritage and the Arts. Cheshire, A., Westphalen, G., Boxall, V., Marsh, R., Gilliland, J., Collings, G., Seddon, S., Loo, M. (2002) Caulerpa taxifolia in West Lakes and the Port River: distribution, eradication options and consequences. South Australian Research and Development Institute, Aquatic Sciences and PIRSA Fisheries, Marine Habitat Program. Publication No. RD02/1061. Chiaramonte, G.E. (1998) The shark genus Carcharhinus Blainville, 1816 (Chondrichthyes: Carcharhinidae) in Argentine waters. Marine and Freshwater Research 49: 747-752. Chick, R.C., Turich, N. and Mayfield, S. (2007) Western Zone Abalone (Haliotis laevigata and H. rubra) Fishery. (Region B). Fishery Status Report to PIRSA. SARDI Aquatic Sciences Publication No.F2007/000563-1. SARDI Research Report Series No.240. SARDI Aquatic Sciences, South Australia. Cliff, G. and Dudley, S.F.J. (1992a) Sharks caught in the protective gill nets off Natal, South Africa. 6. The copper shark Carcharhinus brachyurus (Günther). South African Journal of Marine Science 12: 663–674. Cliff, G. and Dudley, S.F.J. (1992b) Protection against shark attack in South Africa, 1952-90. Australian Journal of Marine and Freshwater Research 43: 263-272 Clifton, J., Olejnik, M., Boruff, B. and Tonts, M. (2007) Patterns of future development in the South-West Marine Region. Institute for Regional Development report, University of Western Australia. Commonwealth of Australia and Government of South Australia (2004) Draft Great Australian Bight Marine Park (Commonwealth Waters and State Waters) Management Plan. Commonwealth of Australia and Government of South Australia. Compagno, L.J. (1984) Sharks of the World. An annotated and illustrated catalogue of the shark species known to date. Volume 1. (Hexanchiformes, Squaliformes, Squatiniformes and Pristiophoriformes). Volume 2. Bullhead, mackerel and carpet sharks (Heterodontiformes, Lamniformes and Orectolobiformes. FAO Species Catalogue for Fisheries Purposes No. 1, Vol.1 and Vol. 2. FAO, Rome. Compagno, L. (2001) FAO Species Catalogue Volume 4. Sharks of the World. Part 1: Hexanchiformes to Lamniformes. Part 2: Carcharhiniformes. FAO Fisheries Synopsis 125/4(1): 1-655. FAO, Rome. Compagno, L.J. (2005) Sharks of the World. An annotated and illustrated catalogue of the shark species known to date. Volume 2. Bullhead, mackerel and carpet sharks (Heterodontiformes, Lamniformes and Orectolobiformes) FAO Species Catalogue for Fisheries Purposes No. 1, Vol.2. FAO, Rome. Compagno, L.J., Ebert, D.A. and Smale, M.J. (1989) Guide to the Sharks and Rays of Southern Africa. New Holland (Publ.) Ltd., London. 158p. Conolly, J.R. and Von Der Borch, C.C. (1967) Sedimentation and physiography of the sea floor south of Australia. Sedimentary Geology 1: 181-187, 191-220 Cotton, B. and Godfrey, F. (1931 and 1932) South Australian shells (including descriptions of new genera and species). South Australian Naturalist 13: 5-23, 35-86 and 127-176. 115 Cotton, B. and Godfrey, F. (1938) The Molluscs of South Australia, Part 1. Pelecypoda. Government Printer, Adelaide, South Australia. 314p. Coulson, P.G., Potter, I.C., Hesp, S.A., and Hall, N.G. (2007) Biological parameters required for managing Western Blue Groper, Blue Morwong and Yellowtail Flathead. Fisheries Research and Development Corporation project report 2004/057. Centre for Fish and Fisheries Research Murdoch University, Murdoch, Western Australia. Coutin, P., Bruce, B.D. and Paul, L.J. (1992) New Zealand school sharks cross the Tasman Sea. Australian Fisheries 51(3): 24-25. Cowan, R. (2006) Australian Marine Algal Name Index. Database by R. Cowan, Murdoch University, Western Australia, for the Australian Biological Resources Study, Canberra. Cowling, A., Polacheck, T. and Miller, C. (1996) Data analysis of the aerial survey (19911996) for juvenile southern bluefin tuna in the Great Australian Bight. Commission for the Conservation of Southern Bluefin Tuna Scientific Committee Meeting CCSBT/SC/96/21. Crawford, C.M., Edgar, G.J. and Cresswell, G. (2000) The Tasmanian region. In: Shepherd, C. and Zan, L.P. (Eds) Seas at the Millennium. Pergamon, Amsterdam. pp. 647–660 CSIRO (1979) Courageous Cruise 047 Biological Data overview, 1979. CSIRO Marine Research, Hobart. CSIRO (1999) 'Lost' fish found 85 years later. Media release: Monday, 26 April, 1999. CSIRO, Hobart. CSIRO (2005) Tagging of White Sharks: Four tagged white sharks (Rolf, Bomber, Michael, Sam C). http://www.cmar.csiro.au/tagging/whitesharks/tracks/michael.html CSIRO (2008a) White Sharks: "Columba" http://www.cmar.csiro.au/tagging/whitesharks/columba/index.htm CSIRO (2008b) White sharks: Bruce and Lulu - first sharks in the cradle. http://www.cmar.csiro.au/tagging/whitesharks/thefirst.html CSIRO (2009) Codes for Australian Aquatic Biota (CAAB). http://www.marine.csiro.au/caab/ Currie, D.R. and Ward, T.M. (2009) South Australian Giant Crab (Pseudocarcinus gigas) Fishery. Fishery Assessment Report for PIRSA. SARDI Research Report Series No. 345. SARDI Aquatic Sciences, South Australia. Currie, D.R., Sorokin, S.J. A and Ward, T.M. (2007) Infaunal assemblages of the eastern Great Australian Bight: Effectiveness of a Benthic Protection Zone in representing regional biodiversity. Final report to the South Australian Department for Environment and Heritage and the Commonwealth Department of the Environment and Water Resources. SARDI Research Report Series No. 250. SARDI Aquatic Sciences, South Australia. Currie, D.R., Sorokin, S.J. A and Ward, T.M. (2008) Performance assessment of the Benthic Protection Zone of the Great Australian Bight Marine Park: epifauna. SARDI Research Report Series No. 299. SARDI Aquatic Sciences, South Australia. 116 Currie, D.R., Sorokin, S.J. A and Ward, T.M. (2009) Infaunal macro-invertebrate assemblages of the eastern Great Australian Bight: effectiveness of a marine protected area in representing the region‘s benthic biodiversity. Marine and Freshwater Research 60 (459–474). Daley, R.K., Stevens, J.D., Last, P.R. and Yearsley, G.K. (2002) Field Guide to Australian Sharks and Rays. CSIRO Marine Research Division, Fisheries Research and Development Corporation and Australian Fisheries Management Authority. 84p. Daley, R. Knuckey, I., Dowdney, J., Williams, A., Bulman, C., Sporcic, M., Fuller, M. and Smith, T. (2006) Draft Ecological Risk Assessment for the Effects of Fishing: Southern and Eastern Scalefish and Shark Fishery - Great Australian Bight Trawl Fishery Sub-fishery. Report for the Australian Fisheries Management Authority, Canberra. Darragh, T.A. (2002) A revision of the Australian genus Umbilia (Gastropoda: Cypraeidae) Memoirs of the Museum of Victoria 59(2): 355–392. Davie, P.J.F. (2001) Crustacea: Malacostraca: In: Australian Biological Resources Study (ABRS) (2009) Australian Faunal Directory. Department of the Environment and Water Resources, Canberra. http://www.environment.gov.au/biodiversity/abrs/online-resources/fauna/afd/index.htm. Davis, R. and Capon, R.J. (1993) Two new scalarane sesterterpenes: isoscalarafuran-A and -B, epimeric alcohols from a Southern Australian marine sponge, Spongia hispida. Australian Journal of Chemistry 46(8): 1295-1299. Dennis, T. and Lashmar, A. (1996) Distribution and abundance of white-bellied sea-eagles in South Australia. Corella 20: 93-102 Dennis, T.E. (2008) Coastal bird and sea lion survey on Alinytjara Wilurara Lands in the Far-west Region of South Australia in 2008. Report to Alinytjara Wilurara NRM Board. September, 2008. Dennis, T.E. and Shaughnessy, P.D. (1996) Status of the Australian sea lion, Neophoca cinerea, in the Great Australian Bight. Wildlife Research 23(6): 741-754. Department for Environment and Heritage (1998) Great Australian Bight Marine Park Management Plan, Part A. Department for Environment, Heritage and Aboriginal Affairs, South Australia. Department for Environment and Heritage (2001) Introduced Pest Guide. Chapter in: Coastcare Community Handbook. pp. 153-157. www.environment.sa.gov.au/coasts/pdfs/coastcare/pests.pdf http://www.environment.sa.gov.au/coasts/pdfs/coastcare/handbook.pdf Department for Environment and Heritage (2003) The Journey of the Southern Right Whale. Brochure. Public Affairs Branch, Department for Environment and Heritage, South Australia. Department for Environment and Heritage (2006) Southern Right Whale - Great Australian Bight Marine Park. Brochure. Department for Environment and Heritage, South Australia, and Australian Government Department of the Environment and Heritage, Canberra. Department for Environment and Heritage (2006) Feral foes: weeds and pest animals. West Coast Babbler: The Ark on Eyre Newsletter. Winter 2006. Department for Environment and Heritage (2006) Island Parks of Western Eyre Peninsula Management Plan. DEH, Adelaide, South Australia. 117 Department for Environment and Heritage (2009) A Review of Nullarbor Regional Reserve 1999 – 2009. Department for Environment and Heritage, South Australia. Department for Environment and Heritage (2009) Marine Park 1 - Far West Coast Marine Park. Department for Environment and Heritage, South Australia. Department for Environment and Heritage (2009) About Marine Parks : Great Australian Bight Marine Park: Legislation and Management. Department for Environment and Heritage, South Australia. Department for Environment and Heritage (2010) Great Australian Bight – Introduction, Location, Legislation and Management. (Web pages). Department of the Environment and Water Resources (2005) Southern Bluefin Tuna (Thunnus maccoyii). Advice to the Minister for the Environment and Heritage from the Threatened Species Scientific Committee (TSSC) on Amendments to the list of Threatened Species under the Environment Protection and Biodiversity Conservation Act 1999 (EPBC Act), 7th September 2005. Department of Environment and Water Resources (2006) A Characterisation of the Marine Environment of the South-west Marine Region. A summary of an expert workshop convened in Perth, Western Australia, September 2006. Department of the Environment, Water, Heritage and the Arts (DEWHA) (2009a) Background paper for the Threat Abatement Plan for the impacts of marine debris on vertebrate marine life. May 2009 Department of the Environment, Water, Heritage and the Arts, Canberra. Department of the Environment, Water, Heritage and the Arts (DEWHA) (2009b) Marine Bioregional Planning in the South-west Marine Region: areas for further assessment. Department of the Environment, Water, Heritage and the Arts, Canberra. Department of the Environment, Water, Heritage and the Arts (DEWHA) (2010) National Whale and Dolphin Sightings and Strandings Database. http://data.aad.gov.au/aadc/whales/ Department of the Environment, Water, Heritage and the Arts (DEWHA) (2010) Eubalaena australis. In: Species Profile and Threats Database. Department of the Environment, Water, Heritage and the Arts, Canberra. http://www.environment.gov.au/sprat Department of Fisheries, Western Australia (2005a) Introduced Marine Aquatic Invaders - a field guide. http://www.fish.wa.gov.au/docs/pub/IMPMarine/IMPMarinePage17.php?0506 Department of Fisheries, Western Australia (2005b) Farming Blue Mussels: Introduction. http://www.fish.wa.gov.au/docs/aqwa/BlueMussels/index.php?0308 Department of Fisheries, Western Australia (2005) Application to the Australian Government Department of Environment and Heritage on the Western Australian Temperate Shark Fisheries (covering the West Coast Demersal Gillnet and Demersal Long-line [interim] Managed Fishery and the Joint Authority Southern Demersal Gillnet and Demersal Long-line Managed Fishery) for consideration under Parts 13 and 13A of the Environment Protection and Biodiversity Conservation Act 1999. Department of Fisheries, Western Australia. Department of Health, South Australia (2009) Toxic algae warning for West Lakes and Port River. Department of Health media release, 28th November 2008. Department of Sustainability and Environment (DSE), Victoria (2009) Action Statement No. 250 118 Flora and Fauna Guarantee Act 1988. Leathery Turtle Dermochelys coriacea. http://www.climatechange.vic.gov.au/CA256F310024B628/0/B6D9D644C34259C7CA2576090019E5A7/ $File/250+Leathery+Turtle+2009.pdf Deveney, M., Rowling, K., Wiltshire, K., Manning, C., Fernandes, M., Collings, G. and Tanner, J. (2008) Caulerpa taxifolia (M. Vahl) C. Agardh: Environmental risk assessment. A report prepared for PIRSA Marine Biosecurity. SARDI Publication No. F2008/000854-1. SARDI Research Report Series No. 307. Dimmlich, W.F., Breed, W.G., Geddes, M. and Ward, T.M. (2004) Relative importance of gulf and shelf waters for spawning and recruitment of Australian anchovy, Engraulis australis, in South Australia. Fisheries Oceanography 13(5): 310–323. Director of National Parks (Australian Government DEH) (2005) Great Australian Bight Marine Park (Commonwealth Waters) Management Plan 2005-2012. Australian Government Department of the Environment and Heritage, Canberra. DITR (2003) Release of offshore petroleum exploration areas, Australia 2003: Ceduna Sub-basin, Bight Basin South Australia. Department of Industry, Tourism and Resources. Dixon, J., Schroeter, S.C. and Kastendiek, J. (1981) Effects of the encrusting bryozoan, Membranipora membranacea, on the loss of blades and fronds by the giant kelp, Macrocystis pyrifera (Laminariales). Journal of Phycology 17(4): 341 – 345. Duffy, C.S. and Gordon, I. (2003). Carcharhinus brachyurus. In: IUCN (2009) IUCN Red List of Threatened Species. www.iucnredlist.org. Edgar, G.J., Samson, C.R. and Barrett, N.S. (2005) Species extinction in the marine environment: Tasmania as a regional example of overlooked losses in biodiversity. Conservation Biology 19: 12941300. Edyvane (1999) Conserving Marine Biodiversity in South Australia – Part 2 – Identification of Areas of High Conservation Value in South Australia, SARDI Report Number 39, PIRSA. Edyvane, K. and Andrews, G. (1998) Great Australian Bight Marine Park Management Plan. Part B: Resource Information. Department for Environment, Heritage and Aboriginal Affairs, South Australia. Edyvane, K. and Baker, J.L. (1998) Marine Benthic Survey of the Eastern Great Australian Bight – Progress Report. Report for the Australian Heritage Commission, National Estate Grants Program. SARDI Aquatic Sciences, South Australia. 27p. Edyvane, K. and Baker, J. (1999) Marine Heritage Values of the Eastern Great Australian Bight, South Australia. Final Report to the Australian Heritage Commission NEGP 1995/96 Project: Survey of the Coastal and Marine Habitats of S.A: Eyre Coast. SARDI Aquatic Sciences, South Australia. Edyvane, K., Dalgetty, A., Hone, P. Higham, J. and Wace, N.M. (2004) Long-term marine litter monitoring in the remote Great Australian Bight, South Australia. Marine Pollution Bulletin 48: 1060-1075. Eveson, P., Farley, J., Bravington, M. and Basson, M. (2007) Southern bluefin tuna aerial survey in the Great Australian Bight – 2007: Preliminary results of aerial survey and commercial spotting data Final Report to the Department of Agriculture, Fisheries and Forestry. CSIRO Marine Research. Eyre Peninsula Tourism Association (2000) Australia’s Great Road Journey: The Nullarbor. (Electronic brochure). 119 FAO (2000) FAO/SIDP - Species Identification and Data Programme Species Identification Sheet Galeorhinus galeus http://www.fao.org/figis/servlet/FiRefServlet?ds=species&fid=2828 Farmer, B. (2008) Comparisons of the biological and genetic characteristics of Mulloway Argyrosomus japonicus Sciaenidae in different regions of Western Australia, PhD thesis, Murdoch University, Western Australia. Farmer, B.M., French, D.J.W., Potter, I.C., Hesp, S.A. and Hall, N.G. (2005) Determination of biological parameters for managing the fisheries for Mulloway and Silver Trevally in Western Australia. Fisheries Research and Development Corporation Report FRDC Project 2002/004. Centre for Fish and Fisheries Research Murdoch University, Murdoch, Western Australia. Feary, D.A., and James, N.P. (1995) Cenozoic biogenic mounds and buried Miocene(?) barrier reef on a predominantly cool-water carbonate continental margin—Eucla basin, western Great Australian Bight. Geology 23(5): 427-430. Feary, D.A., and James, N.P. (1998) Seismic stratigraphy and geological evolution of the Cenozoic, coolwater Eucla Platform, Great Australian Bight. American Association of Petroleum Geologists Bulletin 82(5A): 792–816. Feary, D.A., Birch, G., Boreen, T., Chudyk, E., Lanyon, R., Petkovic, P., and Shafik, S. (1993) Geological Sampling in the Great Australian Bight. Scientific Post-Cruise Report – RV Rig Seismic Cruise 102, Australian Geological Survey Organisation. Vol. Record 1993/18. Feary, D.A., Hine, A.C. and Malone, M.J. (2000) Great Australian Bight: Cenozoic cool water carbonates. Proceedings of the Ocean Drilling Program: Initial Reports 182: 1-58. Feary, D.A., Hine, A.C., James, N.P. and Malone, M.J. (2004) Ocean Drilling Program Leg 182 synthesis: Exposed secrets of the Great Australian Bight. Proceedings of the Ocean Drilling Program: Scientific Results 182 (Special Issue). 30p. Feary, D.A., Hine, A.C., Malone, M.J. and Baldauf, J. (1999) Ocean Drilling Program Leg 182 preliminary report: Great Australian Bight: Cenozoic cool-water carbonates. Preliminary Report of the Ocean Drilling Program, February, 1999. Fergusson, I.K. and Compagno, L. (1995) Annotated elasmobranch species list and bibliography for the Mediterranean. Unpublished report for the IUCN, Gland, Switzerland. Ferguson, G.J., Ward, T.M. and Geddes, M.C. (2008) Do recent age structures and historical catches of mulloway, Argyrosomus japonicus (Sciaenidae), reflect freshwater inflows in the remnant estuary of the Murray River, South Australia? Aquatic Living Resources 21: 145–152. Figueira, W.F. and Booth, D.J. (2009) Increasing ocean temperatures allow tropical fishes to survive over winter in temperate waters. Global Change Biology. 10.1111/j.1365-2486.2009.01934.x Fletcher W.J. (1990) A synopsis of the biology and the exploitation of the Australian pilchard, Sardinops neopilchardus (Steindachner). Part 1: Biology. Fisheries research report no. 88. Fisheries Department of Western Australia, Perth, W.A.. Fowler, A.J., Steer, M., McGarvey, R., and Feenstra, J. (2007) The South Australian Marine Scalefish Fishery Stock Status report. Report to PIRSA. SARDI Research Report Series No. 255. SARDI – Aquatic Sciences, South Australia. 120 Fowler, A.J., McGarvey, R., Steer, M. and Feenstra, J. (2008) The South Australian Marine Scalefish Fishery Stock Status report. Report to PIRSA. SARDI Research Report Series No. 321. SARDI – Aquatic Sciences, South Australia. Fugro (2008) Airborne Magnetic Survey, Great Australian Bight. Environment Plan: Public Summary http://www.pir.sa.gov.au/__data/assets/pdf_file/0013/92200/Fugro_Airborne_Survey_EP__Public_Summary.pdf Gardner, S., Tonts, M. and Elrick, C. (2006) A Socio-economic Analysis and Description of the Marine Industries of Australia’s South-west Marine Region. Final Report Submitted May 2006. Prepared for the Department of the Environment and Water Resources. Gaughan, D., Chidlow, J. and Griffiths, D. (2005) Demersal Gillnet and Long-line Fisheries Status Report. In: Penn, J. (2005) (Ed). WA Fisheries State of the Fisheries Report (2003/04). pp. 186 – 191. Geoscience Australia (2004, 2010), Oil and Gas: Bight Basin. http://www.ga.gov.au/oceans/projects/ssw_bb.jsp Geoscience Australia, CSIRO and Department of the Environment and Heritage (undated) 2005 National Marine Bioregionalisation of Australia: PB32 – Great Australian Bight IMCRA Transition. (Information Sheet) http://www.environment.gov.au/coasts/mbp/publications/general/pubs/pb32.pdf Gibbs, S. (2004) Resource overlap in dolphins and fisheries using Spencer Gulf, South Australia. (Presentation). South Australian Museum. Goldsworthy, S.D. (2008) FRDC PN 2007/041. Mitigating seal interactions in the SRLF and the Gillnet Sector SESSF in South Australia. Milestone report No. 1 to FRDC, May, 2008. SARDI – Aquatic Sciences, South Australia. Goldsworthy, S.D. and Page, B. (2007) a risk-assessment approach to evaluating the significance of seal bycatch in two Australian fisheries. Biological Conservation 139: 269-285. Goldsworthy, S.D. and Page, B. (2009) A review of the distribution of seals in South Australia. SARDI – Aquatic Sciences Publication No. F2009/000368-1. SARDI – Aquatic Sciences, South Australia. Goldsworthy, S., Campbell, R. and McKenzie, J. (2007) Pinnipeds. Chapter in: McClatchie, S. et al. (Eds) Review of Ecological Information and Knowledge of Australia‘s South West Marine Region (SWMR). Report by SARDI Aquatic Sciences and University of Western Australia, for National Oceans Office, Canberra. Goldsworthy, S.D., McKenzie, J., Shaughnessy, P.D., Macintosh, R.R., Page, B., and Campbell, R. (2009a) An update of the report: Understanding the Impediments to the Growth of Australian Sea Lion Populations. Report to the Department of the Environment, Water, Heritage and the Arts. SARDI Publication Number F2008/00847-1. SARDI Research Report series No. 356. 175p. Goldsworthy, S.D., Page, B., Lowther, A., Shaughnessy, P.D., Peters, K.P., Rogers, P., McKenzie, J., Bradshaw, C.J.A. (2009b) Developing population protocols to determine the abundance of Australian sea lions at key subpopulations in South Australia. Final Report to the Australian Marine mammal Centre, Department of the Environment, Water, Heritage, and the Arts. SARDI Aquatic Sciences Publication Number F2009/000161-1, SARDI Research Report Series No. 348. 58p. Government of Western Australia (2006) Marine Futures: Site Selection Workshop Report. 31p. Gowlett-Holmes, K. (2008) A Field Guide to the Marine Invertebrates of South Australia. Notomares, Sandy Bay, Tasmania. 121 Guiry, M.D. and Guiry, G.M. (2010) AlgaeBase. World-wide electronic publication. National University of Ireland, Galway. http://www.algaebase.org. Gunasekera, R.M., Patil, J.G., McEnnulty, F. R. and Bax, N.J. (2005) Specific amplification of mt-COI gene of the invasive gastropod Maoricolpus roseus in planktonic samples reveals a free-living larval lifehistory stage. Marine and Freshwater Research 56(6): 901–912. Gunn, J. and Young, J. (1999) Environmental determinants of the movement and migration of juvenile southern bluefin tuna. In: Hancock, D.A., Smith, D.C., and Koehn, J.D (Eds) Fish Movement and Migration. Australian Society for Fish Biology Workshop Proceedings, Bendigo, Victoria, September 1999. Australian Society for Fish Biology, Sydney. pp. 123-128 Gurgel, C. F. (2009) Status of the genus Sargassum in the Great Australian Bight (AW NRM). Plant Biodiversity Centre report, State Herbarium of South Australia. Halpern, B., Selkoe, K.A., Micheli F., and Kappel, C.V. (2007) Evaluating and ranking the vulnerability of global marine ecosystems to anthropogenic threats. Conservation Biology 21: 1301-1315. Hamer, D.J., Ward, T.M., Goldsworthy, S.D., and Shaughnessy P.D. (2009) Effectiveness of the Great Australian Bight Marine Park in protecting the Australian sea lion (Neophoca cinerea) from by-catch mortality in shark gill-nets. Report to Great Australian Bight Marine Park Steering Committee. SARDI Aquatic Sciences Publication No. F2009/000227-1, SARDI Research Report Series No. 357. 59pp. Harrison, N. (2001) A Five-Year Management Strategy for Recreational Fishing on the West Coast of Western Australia. Final report of the West Coast Recreational Fishing Working Group, compiled by N. Harrison, with advice from the West Coast Recreational Fishing Working Group. August 2001. Fisheries Management Paper No. 153. W.A. Fisheries. 114p. Hayes, K., Sliwa, C., Migus, S. McEnnulty, F. and Dunstan, P. (2005) National priority pests: Part II. Ranking of Australian marine pests. An independent report undertaken for the Department of Environment and Heritage by CSIRO Marine Research, Hobart. Henning, T. and Hemmen, J. (1993) Ranellidae and Personidae of the World. V.C. Hemmen, Wiesbaden, Germany. 263p. henrythesealion.com (2010) (i) Lily the Amazing Mako shark; (ii) Shark Research in South Australia http://www.henrythesealion.com/lilly/tracking/index.html http://www.henrythesealion.com/lilly/research/index.html Hewish, D. and Gowlett-Holmes, K. (1991). Mollusc type specimens in the South Australian Museum. Rec. South Aust. Museum 25: 57-70. Herzfeld, M. (1996) Sea surface temperatures and circulation in the Great Australian Bight. Ph.D. Thesis, School of Earth Sciences, Faculty of Science and Engineering, Flinders University of South Australia. Hertzfeld, M. (1997) The annual cycle of sea-surface temperature in the Great Australian Bight. Progress in Oceanography 39: 1-27. Herzfeld, M. (2000) Water mass formation, circulation and productivity in the Great Australian Bight. Seminar at CSIRO Auditorium, Friday, 22 June, 2000, 11.15 am. CSIRO Marine Research, Hobart. http://www.marine.csiro.au/seminars/sem-abs00/herzfeld.html Herzfeld, M. and Tomczac, M. (1997) Bottom-driven upwelling generated by eastern intensification in closed and semi-enclosed basins with a sloping bottom. Marine and Freshwater Research 50: 613-627. 122 Hertzfeld, M., and Tomczak, M. (1999) Bottom driven upwelling generated by eastern intensification in closed and semi-closed basins. Marine and Freshwater Research 50: 613-627. Hill, P.J., Rollet, N., and Symonds, P. (2001) Seafloor mapping of the South-east Marine Region and adjacent waters - AUSTREA final report: Lord Howe Island, south-east Australian margin (includes Tasmania and South Tasman Rise) and central Great Australian Bight. Australian Geological Survey Organization Record 2001/08, 83p. Hine, A.C., Feary, D.A., Malone, M.J. et al. (1999) Research in Great Australian Bight yields exciting early results. Eos, Transactions of the American Geophysical Union 80(44): 521, 525-526. Hobday, A.J. (2002) Acoustic monitoring of SBT within the GAB: Exchange rates and residence times at topographic features. CSIRO Marine Research report RMWS/02/03. CSIRO, Hobart. Hobday, A.J., Okey, T.A., Poloczanska, E.S., Kunz, T.J. and Richardson, A.J. (Eds) (2006) Impacts of climate change on Australian marine life. Part A. Executive Summary. Part B: Technical Report. Part C: Literature Review. Reports to the Australian Greenhouse Office, Canberra, Australia. September 2006. Hobday A.J., Griffiths S. and Ward T. (2009) Pelagic fishes and sharks. In: Poloczanska, E.S., Hobday, A.J. and Richardson, A.J. (Eds) A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009. NCCARF Publication 05/09, ISBN 978-1-921609-03-9. Holbrook, N.J., Davidson, J., Feng, M., Hobday, A.J., Lough, J.M., McGregor, S. and Risbey, J.S. (2009) El Niño-Southern Oscillation. In: Poloczanska, E.S., Hobday, A.J. and Richardson, A.J. (Eds) A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009. NCCARF Publication 05/09, ISBN 978-1-921609-03-9. Hunt, S. (2006) Fowlers Bay offshore (3/2006). (Fishing article). http://www.marinews.com/fishing_details.php?recordid=382 Hurst, R.J., Bagley, N.W., McGregor, G. and Francis, M.P. (1999a) Movements of New Zealand school shark, Galeorhinus galeus, from tag returns. New Zealand Journal of Marine and Freshwater Research 33: 29-48. Hurst, R.J., Bagley, N.W., McGregor, G. and Francis, M.P. (1999b) New Zealand school sharks – our loss is Australia‘s gain? Fishing Today 12(1): 36-37. Hutchins, J.B., and Swainston, R. (1986 and 2001) Sea Fishes of Southern Australia. Swainston Publishing, Perth, Western Australia. Huveneers, C. (2006) Redescription of two species of wobbegongs (Chondrichthyes: Orectolobidae) with elevation of Orectolobus halei Whitley 1940 to species level. Zootaxa 1284: 29-51. Huveneers, C., Harcourt, R. and Otway, N. (2006) Observation of localised movements and residence times of the wobbegong shark Orectolobus halei. Cybium 30(4) (Supplement): 103-111. Huveneers, C. (2007) The ecology and biology of wobbegong sharks (genus Orectolobus) in relation to the commercial fishery In New South Wales, Australia. PhD thesis. Macquarie University, New South Wales. IMCRA Technical Group (1998) Interim Marine and Coastal Regionalisation for Australia: an ecosystembased classification for marine and coastal environments, Version 3.3. Australian Government, Canberra. 123 Izzo, C. (2006) Aging Port Jackson sharks, Heterodontus portusjacksoni (Myer 1793), using calcified structures. Honours Thesis, Flinders University, Adelaide. 105p. James, N.P. and Von der Borch, C.C. (1991) Carbonate shelf edge off southern Australia: a prograding open-platform margin. Geology 19: 1005-1008. James, N.P., Boreen, T. Bone, Y. and Feary, D. (1994) Holocene carbonate sedimentation on the west Eucla Shelf, Great Australian Bight: a shaved shelf. Sedimentary Geology 90: 161-177. James, N.P., Bone, Y. and Hageman, S.J. (1996) The Lioness, Bonaparte's Tongue and cool-water carbonates on the South Australian continental shelf. AAPG Bulletin 80. James, N.P., Bone, Y., Collins, L.B. and Kyser, T.K. (2001) Surficial sediments of the Great Australian Bight: facies dynamics and oceanography on a vast cool-water carbonate shelf. Journal of Sedimentary Research 71(4): 549-567. James, N.P., Boreen, T.D., Bone, Y. and Feary, D.A. (1994) Holocene carbonate sedimentation on the west Eucla Shelf, Great Australian Bight: a shaved shelf. Sedimentary Geology 90 (3-4): 161-177. James, N.P., Feary, D.A., Betzler, C., Bone, Y., Holbourn, A.E., Li, Q., Machiyama, H., Toni Simo, J.A. and Surly, F. (2004) Origin of late Pleistocene bryozoan reef mounds: Great Australian Bight. Journal of Sedimentary Research 74(1): 20–48. James, N.P., Feary, D.A., Surlyk, F., Simo, J.A.T., Betzler, C., Holbourn, A.E., Li, Q., Matsuda, H., Machiyama, H., Brooks, G.R., Andres, M.S., Hine, A.C., Malone, M.J. and Ocean Drilling Program Leg 182 Scientific Party (2000) Quaternary bryozoan reef mounds in cool-water, upper slope environments: Great Australian Bight. Geology 28(7): 647-650. Jones, D. and Morgan, G. (2002) A Field Guide to Crustaceans of Australian Waters. Reed New Holland Publishers, Sydney, Australia. 224p. Jones, G.K. (1991) Fin fishery considerations in the management of the proposed Great Australian Bight Marine Park. Safish 15(4): 11. Jones, G.K. (2008) Review of the Fishery for Whaler Sharks (Carcharhinus spp.) in South Australian waters. SARDI Aquatic Sciences Publication F2007/00721-1. SARDI Aquatic Sciences, South Australia. Jones, G.K. (2009) South Australian Recreational Fishing Survey. PIRSA Fisheries, Adelaide, 84p. South Australian Fisheries Management Series Paper No 54. Jones, G.K. and Doonan, A. (2005) 2000/1 National Recreational and Indigenous Fishing Survey: South Australian regional information. Department of Primary Industries and Resources, Adelaide, South Australia. Jones, M. (1995) Fishing debris in the Australian marine environment. Marine Pollution Bulletin 30(1): 25-33. Joung, S-J., Liao, Y-Y., Chou, Y-C., Liu, H-C. and Chen, C-T. (2005) Age, growth, and reproduction of smooth hammerhead, Sphyrna zygaena in north-eastern Taiwan waters. Paper presented at the 7th Indo-Pacific Fish Conference, 16- 21 May, 2005, Taipei, Taiwan. Kailola, P.J. and Jones, G.K. (1981) First record of Promicrops lanceolatus (Bloch) (Pisces: Serrandiae) in South Australian waters. Transactions of the Royal Society of South Australia 105 (4): 211-212. 124 Kailola, P., Williams, M., Stewart, P., Reichelt, R., McNee, A. and Grieve, C. (1993) Australian Fisheries Resources. Bureau of Resource Sciences, and FRDC, Canberra. Kemper, C.M. (1994) Possible influences of oceanographic features of GAB on cetaceans. In: Andrews, G. (Ed.) The Great Australian Bight Marine Park, Ceduna Workshop. Proceedings and discussion papers. SARDI Aquatic Sciences, South Australia. Kemper, C.M. (1998) Cetaceans in the Great Australian Bight. In: Slater, J (Ed.) Working Draft. Workshop to identify research and monitoring priorities for the Great Australian Bight Marine Park. Environment Australia. Technical Report No. 1. Kemper, C. (2004) Cetacean Interactions with Fisheries and Aquaculture. (Presentation). South Australian Museum. Kemper, C. and Gibbs, S. (2001) Dolphin interactions with tuna feedlots at Port Lincoln, South Australia and recommendations for minimising entanglements. Journal of Cetacean Research and Management 3: 283−292. Kemper, C., Pemberton, D., Cawthorn, M., Heinrich, S., Mann, J., Wursig, B., Shaughnessy, P. and Gales, R. (2003) Aquaculture and marine mammals: co-existence or conflict? In: Gales, N., Hindell, M. and Kirkwood, R. (Eds) Marine Mammals and Humans: Fisheries, Tourism and Management Issues. CSIRO Publishing, Melbourne, Victoria. Kemper, C. and Ling, J. (1991) Whale strandings in South Australia (1881-1989). Transactions of the Royal Society of South Australia 115(1): 37-52. Kemper, C.M., Flaherty, A., Gibbs, S.E., Hill, M., Long, M. and Byard, R.W. (2005) Cetacean captures, strandings and mortalities in South Australia 1881−2000, with special reference to human interactions. Australian Mammalogy 27: 37-47. Klaer, N.L. (2001) Steam trawl catches from south-eastern Australia from 1918 to 1957: trends in catch rates and species composition. Marine and Freshwater Research 52: 399–410. Klaer, N.L. (2007) Updated stock assessment for deepwater flathead (Neoplatycephalus conatus) and Bight redfish (Centroberyx gerrardi) in the Great Australian Bight trawl fishery using data to June 2007. CSIRO, Hobart. Klimley, A.P. and Ainley, D.G. (Eds) (1996) Great White Sharks. The biology of Carcharodon carcharias. Academic Press, New York. Kloser, R., Ryan, T., Sakov, P., Young, J.D., and Davis, T. (1998) Assessment of acoustics as a research tool to determine the distribution, biomass and behaviour of southern bluefin tuna schools and their prey in the Great Australian Bight. In: Report of Tenth Workshop, 14–17th September. Hobart, Tasmania: CSIRO. Knight, M., Tsolos, A. and Doonan, A. (2006) South Australian Fisheries and Aquaculture Information and Statistics Report 2006. SARDI Research Report Series No 118. SARDI Aquatic Sciences, South Australia. Knight, M., Doonan, A. and Tsolos, A. (2007) South Australian Wild Fisheries Information and Statistics Report. SARDI Research Report Series No 200. SARDI Aquatic Sciences, South Australia. 125 Knight, M., Doonan, A. and Tsolos, A. (2007) The South Australian Recreational Charter Boat Fishery. SARDI Aquatic Sciences Publication No. F2007/000847-1. SARDI Research Report Series No. 239, September 2007. Knight, M. and Tsolos, A. (2009) South Australian Wild Fisheries Information and Statistics Report. SARDI Research Report Series No. 305. January, 2009. Knox, G.A. (1963) The biogeography and intertidal ecology of the Australasian coasts. Oceanography and Marine Biology: An Annual Review. 1: 341-404. Kott, P. (1972) The ascidians of South Australia II. Eastern Sector of the Great Australian Bight and Investigator Strait. Transactions of the Royal Society of South Australia 96(4):165-196. Kott, P. (1975) The ascidians of South Australia III. Northern sector of the Great Australian Bight and additional records. Transactions of the Royal Society of South Australia 99(1): 1-20. Knuckey, I., Koopman, M., and Hudson, R. (2009). Resource Survey of the Great Australian Bight Trawl Fishery 2009. AFMA Project 2008/848. Fishwell Consulting 20p. Kuiter, R.H. (1996a) Guide to Sea Fishes of Australia. (New Holland Publishers Australia Pty Ltd). 430p. Kuiter, R.H. (1996b) The Complete Divers’ Guide to Coastal Fishes of South-Eastern Australia. Natural Learning Titles, CD-ROM. Kyne, P.M., Cavanagh, R., Fowler, S.L. and Pollock, C. (2005) IUCN Shark Specialist Group Red List assessments, 2000 - 2005. Table produced by IUCN Shark Specialist Group, IUCN, Switzerland. Kyne, P.M., Cavanagh, R., Fowler, S.L. and Pollock, C. (2006) IUCN Shark Specialist Group Red List assessments, 2000 - 2006. Table produced by IUCN Shark Specialist Group, IUCN, Switzerland. Kyne, P.M., Johnson, J.W., Courntney, A. and Bennett, M.B. (2005) New biogeographical information on Queensland chondrichthyans. Memoirs of the Queensland Museum 50(2): 321-327. Larcombe, J. and Begg, G. (2008) Fishery Status Reports 2007. Status of fish stocks managed by the Australian Government. Bureau of Rural Sciences, Canberra. Last, P.R. and Stevens, J.D. (1994) Sharks and Rays of Australia. CSIRO Publishing, Australia. Last, P.R. White, W.T., Pogonoski, J.J., and Gledhill, D.C. (2008) Descriptions of new Australian skates (Batoidea: Rajoidei). CSIRO Marine and Atmospheric Research Paper; 021. CSIRO, Hobart. Lenanton, R.C., Heald, D.I., Platell, M., Cliff, M. and Shaw, J. (1990) Aspects of the reproductive biology of the gummy shark, Mustelus antarcticus Günther, from waters off the south coast of Western Australia. Australian Journal of Marine and Freshwater Research 41: 807–822. Lenanton, R.C., Joll, L., Penn, J.W. and Jones, K. (1991) The influence of the Leeuwin Current on coastal fisheries in Western Australia. Journal of the Royal Society of Western Australia 74: 101–114. Lindholm, R. (1984) Observations on the chinaman leatherjacket Nelusetta ayraudi (Quoy & Gaimard) in the Great Australian Bight. Australian Journal of Marine and Freshwater Research 35(5): 597-99. Ling, J. (1991). Recent sightings of killer whales Orcinus orca (Cetacea: Delphinidae) in South Australia. Transactions of the Royal Society of South Australia 115(2): 95-98. 126 Linnane, A., McGarvey, R. and Feenstra, J. (2008) Northern Zone Rock Lobster Fishery 2006/ 2007. SARDI Aquatic Sciences Publication No. F2007/000320-2. SARDI Aquatic Sciences, South Australia. Linnane, A., McGarvey, R. and Feenstra, J. (2008) Northern Zone Rock Lobster Fishery Status Report 2007/ 2008. SARDI Aquatic Sciences Publication No. F2007/000714-2. SARDI Aquatic Sciences, South Australia. Linnane, A., McGarvey, R. and Feenstra, J. (2009) Northern Zone Rock Lobster Fishery 2007/ 2008. SARDI Aquatic Sciences Publication No. F2007/000320-3. SARDI Aquatic Sciences, South Australia. Li, Q. and McGowran, B. (1998) Oceanographic implications of recent planktonic foraminifera along the southern Australian margin. Marine and Freshwater Research 49(5): 439-445. Li, Q., McGowran, B., and James, N.P. (2003) Eocene–Oligocene planktonic forminiferal biostratigraphy of Sites 1126, 1130, 1132, and 1134, ODP Leg 182, Great Australian Bight. In: Hine, A.C., Feary, D.A., and Malone, M.J. (Eds.) Proc. ODP, Sci. Results 182: 1–28. Ling J.K. and Needham, D.J. (1985) Southern right whale aerial surveys, South Australia. Annual reports 1985-91. Australian National Parks and Wildlife Service, Canberra. (Unpublished report). Ling, J.K. and Needham, D.J. (1994) Incidental sightings and aerial observations of southern right whales Eubalaena australis in South Australia. Unpublished report presented at the Great Australian Bight Marine Park Workshop. Ceduna, 1994. Edited by G.J. Andrews. SARDI, West Beach, South Australia. Ling, S. (2008) Range expansion of a habitat-modifying species leads to loss of taxonomic diversity: a new and impoverished reef state. Oecologia 156: 883-894. Ling, S.D. and Johnson, C.R. (2009) Population dynamics of an ecologically important range-extender: kelp beds versus sea urchin barrens. Marine Ecology Progress Series 374: 113-125. Ling, S.D., Johnson, C.R., Frusher, S. and King, C.K. (2008) Reproductive potential of a marine ecosystem engineer at the edge of a newly expanded range. Global Change Biology 14: 907-915. Ling, S.D., Johnson, C.R., Ridgway, K., Hobday, A.J. and Haddon, M. (2009) Climate-driven range extension of a sea urchin: inferring future trends by analysis of recent population dynamics. Global Change Biology 15: 719-731. Lisk, M., Hall, D., Ostby, J. and Brincat, M.P. (2001) Addressing the oil migration risks in the Great Australian Bight. In: Hill K.C. and Bernecker, T. (Eds) Eastern Australasian basins symposium 2001. A refocused energy perspective for the future. Petroleum Exploration Society of Australia Special Publication 1: 553-562. Lowry, D.C. (1970) Geology of the western Australian part of the Eucla Basin. Western Australia Geological Survey Bulletin 122. 1-201. Lowry, D.C. and Jennings, J.N. (1974) The Nullarbor Karst, Australia. Zeitschrift für Geomorphologie 18(1): 36-81. Lynch, A.W. and Garvey, J.R. (2003) Great Australian Bight Trawl Fishery Data Summary 1990-2002. Logbook Program. Australian Fisheries Management Authority, Canberra. Markina, N.P. and Vishnyakova, O.V. (1977) Species composition and quantitative distribution of phytoplankton from the Great Australian Bight. Sov. J. Mar. Biol. 3(2): 96-102. 127 Martin, S. and Gattuso, J.P. (2009) Response of Mediterranean coralline algae to ocean acidification and elevated temperature. Global Change Biology 15:2089-2100. Mawson, P.R. and Coughran, D.K. (1999) Records of sick, injured and dead pinnipeds in Western Australia 1980-1996. Journal of the Royal Society of Western Australia 82: 121-128. Maxwell, G.J.H. and Cresswell, G.R. (1981) Dispersal of tropical marine fauna to the Great Australian Bight by the Leeuwin Current. Australian Journal of Marine and Freshwater Research 32 (4): 493-500. McAuley, R. (2005) Appendix 6.3: Status Report for the Southern and West Coast Demersal and Demersal Longline Fisheries and Northern Shark Fisheries. Report for Department of the Environment, Water, Heritage and the Arts. Department of Fisheries, Western Australia. http://www.environment.gov.au/coasts/fisheries/wa/temperate-shark/pubs/wa-temperate-shark-appendix63.pdf McCauley, R. (2004) Marine biological acoustics / Underwater noise sources / Environmental effects of noise / Seabed mapping with active systems. (Presentation)/Centre for Marine Science and Technology, Curtin University. McCauley R.D., Fewtrell, J., Duncan, A.J., Jenner, C., Jenner, M-N., Penrose, J.D., Prince, R.I.T., Adhitya, A., Murdoch J. and McCabe, K. (2000) Marine seismic surveys - A study of environmental implications. APPEA Journal 2000: 692-708. McClatchie, S. (1979) Grazing of Zeacumantus subcarinatus Gastropoda on Ulva lactuca. Mauri Ora 7: 39-46. McClatchie, S., Middleton, J.F. and Ward, T.M. (2006) Water mass analysis and alongshore variation in upwelling intensity in the eastern Great Australian Bight. Journal of Geophysical Research 11: C08007.1 - C08007.13. McClatchie, S., Middleton, J., and Kendrick, G. (2007) Ecological integration: links to ocean circulation processes. Chapter 3.5 in: McClatchie, S., Middleton, J., Pattiaratchi, C., Currie, D. and Kendrick, G. (Eds) The South-west Marine Region: Ecosystems and Key Species Groups. Report by SARDI Aquatic Sciences and University of Western Australia, for National Oceans Office, Canberra. McClatchie, S., Middleton, J., Pattiaratchi, C., Currie, D. and Kendrick, G. (Eds) The South-west Marine Region: Ecosystems and Key Species Groups. Report by SARDI Aquatic Sciences and University of Western Australia, for National Oceans Office, Canberra. McEnnulty, F.R., Jones, T.E. and Bax, N.J. (2001) The Web-Based Rapid Response Toolbox. http://crimp.marine.csiro.au/NIMPIS/controls.htm. Date of release: June 2001. McLaughlin, R.H. and O'Gower, A.K. (1971) Life history and underwater studies of a Heterodont shark. Ecological Monographs 41: 271-289. McLeay, L.J. Sorokin, S.J., Rogers, P.J. and Ward, T.M. (2003) Benthic Protection Zone of the Great Australian Bight Marine Park: 1. Literature Review. Final report to National Parks and Wildlife South Australia and the Commonwealth Department of the Environment and Heritage. December, 2003. South Australian Research and Development Institute (Aquatic Sciences), South Australia. McMinn, A., Hallegraeff, G., Smith, J., Lovell, A., Jenkinson, A. and Heijnis, H. (2000) Recent appearance of Gymnodinium catenatum at Port Lincoln, South Australia? Presented at the 9th International Conference on Algal Blooms, Tasmania. 128 McNeil, B.I. and Matear, R.J. (2008) Southern Ocean acidification: A tipping point at 450-ppm atmospheric CO2. Proceedings of the National Academy of Science 105: 18860-18864. Moore, S.K., Trainer, V.L., Mantua, N.J., Parker, M.S., Laws, E.A., Backer, L.C. and Fleming, L.E. (2008) Impacts of climate variability and future climate change on harmful algal blooms and human health. Environmental Health 7 (Suppl 2): S4. Morison, A., Patterson, H. and Pham, T. (2009) Great Australian Bight Trawl Sector. Chapter 11 in: Wilson, D., Curtotti, R., Begg, G. and Phillips, K. (Eds) (2009) Fishery Status Reports 2008: Status Of Fish Stocks And Fisheries Managed By The Australian Government. Bureau of Rural Sciences & Australian Bureau of Agricultural and Resource Economics, Canberra. Morrice, M. (undated) Killer whales (Orcinus orca) in Australian territorial waters. http://www.environment.gov.au/coasts/species/cetaceans/conference/pubs/research-morrice.pdf Moulton, P.M., Walker, T.I. and Saddlier, S.R. (1992) Age and growth studies of gummy shark, Mustelus antarcticus (Günther), and school shark, Galeorhinus galeus (Linnaeus), from southern Australian waters. Australian Journal of Marine and Freshwater Research 43: 1241–1267. Muñoz-Chápuli, R. (1984) Ethologie de la reproduction chez quelques requins de l‘Atlantique nord-est. Cybium 8(3): 1–14. Murton, S. (2003) The complete experience. (Fishing article on Far West Coast fishing tours). South Australian Angler. April / May, 2003. pp.3-6. Murray, L. Lim, T.K., Currie, G. and Capon, R.J. (1995) Aplidites (A-G): Macrocyclic orthonitrites from an Australian tunicate, Aplidium sp. Australian Journal of Chemistry 48(7): 1253-1256. Musick, J.A., Grubbs, R.D., Baum, J. and Cortés, E. (2007) Carcharhinus obscurus. In: IUCN (2009) IUCN Red List of Threatened Species. Version 2009.2. <www.iucnredlist.org>. National Seal Strategy Group and Stewardson, C. (2005) National Assessment of Interactions between Humans and Seals: Fisheries, Aquaculture and Tourism. Report for the Marine and Coastal Committee of the Natural Resource Management Standing Committee. DAFF, Canberra. Newton, G. and Klaer, N. (1991) Deep-sea demersal fisheries resources of the Great Australian Bight: a multi-vessel trawl survey. Bureau of Rural Resources Bulletin No. 10. Australian Government Publishing Service, Canberra. Newton, G. and Klaer, N. (1991) Great Australian Bight Development Trawl Fishery - progress report to GABIA. Bureau of Rural Resources. 8p. Newton, G., Smith, D., Burnell, S., Turner, D. and Robertson, S. (1994) Deepwater Flathead (Neoplatycephalus conatus) and Bight redfish (Centroberyx gerrardi) from the Great Australian Bight. I. Aspects of the Biology (Draft report). NIMPIS (2002) (i) Styela plicata species summary. (ii) Styela clava species summary. (iii) Antennella secundaria species summary. Northern Pacific seastar Asterias amurensis species summary. In: Hewitt C.L., Martin R.B., Sliwa C., McEnnulty, F.R., Murphy, N.E., Jones T. and Cooper, S. (Eds) National Introduced Marine Pest Information System. Web publication <http://crimp.marine.csiro.au/nimpis (URL accessed December 2009). NIMPIS (2009, 2010) National Introduced Marine Pest Information System. http://www.marinepests.gov.au/ 129 Noye, B. J., Matthews, K. and Grzechnik, M.P. (1998) A three-dimensional model of tides and surges in the Great Australian Bight. Modeling Coastal Sea Processes – Ocean and Atmospheric Pacific OA95, Ed. B. J. Noye, World Sci. Pub. Co., Singapore, 1-27. O‘Connor, S. et al. (2004) From whalers to whale watchers - the growth of whale watching tourism in Australia. Report for the International Fund for Animal Welfare, by Economists@Large & Associates, Consulting Economists, Noble Park, Victoria. Ogle, G. (2004) Comments on the Draft Management Plan for The Great Australian Bight Marine Park (Commonwealth Waters). 3rd December 2004. The Wilderness Society, South Australia. O‘Hara, T. (2002) Endemism, rarity and vulnerability of marine species along a temperate coastline Invertebrate Systematics 16(4): 671 – 684. Olsen, A.M. (1954) The biology, migration, and growth rate of the school shark, Galeorhinus australis (Macleay) (Carcharhinidae) in south-eastern Australian waters. Australian Journal of Marine and Freshwater Research 5: 353–410. Olsen, A.M. (1984) Synopsis of biological data on the school shark, Galeorhinus australis (Macleay, 1881). FAO Fisheries Synopsis 139. FAO, Rome. 42p. O‘Neill, B.J. (2004) History of petroleum exploration. In: O‘Brien, G.W. and Hibburt, J. (Eds) Petroleum Geology of South Australia. Volume 5. Great Australian Bight. Department of Primary Industries and Resources, South Australia. Ovenden, S.P.B. and Capon, R.J. (1999) Nuapapuin A and Sigmosceptrellins D and E: New Norterpene Cyclic Peroxides from a Southern Australian Marine Sponge, Sigmosceptrella sp. Journal of Natural Products 62(2): 214-218. Page, B., McKenzie, J., McIntosh, R., Baylis, A., Morrissey, A., Calvert, N., Haase, T., Berris, M., Dowie, D., Shaugnessy, P. and Goldsworthy, S.D. (2004). Entanglement of Australian sea lions and New Zealand fur seals in lost fishing gear and other marine debris before and after Government and industry attempts to reduce the problem. Marine Pollution Bulletin 49: 33–42. Patenaude, N.J. and Harcourt, R. (2006) Australia's southern right whale stock differentiation. Final Report. Australian Government Department of Environment and Heritage, Canberra. (Unpublished report). Pearce A.F. and Caputi N. (1994) Effects of seasonal and inter-annual variability of the ocean environment of recruitment to the fisheries of Western Australia, FRDC Final Report 1994/032. Pearce, A.F. and Feng, M. (2007) Observations of warming in the Western Australian continental shelf. Marine and Freshwater Research 58: 914 – 920. PIRSA (1999) Green lipped mussels in South Australian waters. Pamphlet. Coast and Clean Seas/ NHT. Primary Industries and Resources, South Australia. PIRSA (2007) Introduced Marine Pests of Concern in South Australia. Pamphlet. Primary Industries and Resources, South Australia. Pirzl, R.M. (2004) Habitat preference of Southern Right Whales (Eubalaena australis) in Australian coastal waters. Progress Report to the Department of Environment and Heritage. Deakin University, Warrnambool, Victoria. 130 Pirzl, R.M. (2008) Spatial ecology of Eubalaena australis: habitat selection at multiple scales. Ph.D. thesis. School of Life and Environmental Sciences, Deakin University, Victoria. Pirzl, R.M. and Burnell, S. (2004) A case study of research on southern right whales at Head of Bight, South Australia. (Presentation). Deakin University and Eubalaena Pty Ltd. Pogonoski, J.J., Pollard, D.A. and Paxton, J.R. (2002) Conservation Overview and Action Plan for Australian Threatened and Potentially Threatened Marine and Estuarine Fishes. Environment Australia, Canberra. Poloczanska, E. S., Hobday, A. J. and Richardson, A. J. (Eds) (2008) In Hot Water: preparing for climate change in Australia’s coastal and marine systems. Proceedings of conference held in Brisbane, 12-14th November 2007, CSIRO Marine and Atmospheric Research, Hobart, Tasmania, Australia. Ponder, W. and Grayson, J. (1998) The Australian Marine Molluscs Considered to be Potentially Vulnerable to the Shell Trade. Australian Museum Report, for Environment Australia, Canberra. Ponder, W. and de Keyzer, R. (1992) A revision of the genus Diala (Gastropoda; Cerithioidea; Dialidae). Invertebrate Taxonomy 6:1019-1075. Ponder, W. and Yoo, E. (1976). A revision of the Australian and tropical Indo-Pacific Tertiary and Recent species of Pisinna (=Estea) (Mollusca: Gastropoda: Rissoidae). Records of the Australian Museum 30(10):150–247. Ponder, W. and Yoo, E. (1977a). Revision of Australian species of Rissoellidae (Mollusca). Records of the Australian Museum 31(4):133–85. Figures 1–14. Ponder, W. and Yoo, E. (1977b). A revision of the Eatoniellidae of Australia (Mollusca, Gastropoda, Littorinacea). Records of the Australian Museum 31(15):606–58. Ponder, W. and Yoo, E. (1980). A review of the genera of the Cingulopsidae, with a revision of the Australian and tropical Indo-Pacific species. Records of the Australian Museum 33(1):1-88. Ponder, W., Hutchings, P. and Chapman, R. (2002). Overview of the Conservation of Australian Marine Invertebrates. Prepared by Australian Museum for Environment Australia, Canberra. Poore, G.C.B. (2004) Marine Decapod Crustacea of Southern Australia. A Guide to Identification. CSIRO Publishing, and Museum of Victoria. Powter, D. (2007) Life history and demography of Port Jackson Shark, Heterodontus portusjacksoni. Ph.D. thesis, University of Newcastle, New South Wales. Prince, J.D. (1996) Report on the Southern Shark Fishery pupping workshop held at Queenscliff, August, 1994. FRDC 1996/063. Biospherics Pty Ltd, Leederville, W.A.. Pyle, P., Schramm, M.J., Keiper, C. and Anderson, S.D. (1999) Predation on a white shark (Carcharodon carcharias) by a killer whale (Orcinus orca) and a possible case of competitive displacement. Marine Mammal Science 15: 563–568. Reynolds, S. (2005) Numerous nudibranch findings. Marine Life Society of South Australia Inc. Newsletter No. 322, June 2005. 131 Richardson, L., Mathews, E. and Heap, A. (2005) Geomorphology and Sedimentology of the South Western Planning Area of Australia: review and synthesis of relevant literature in support of Regional Marine Planning. Geoscience Australia, Record 2005/17. 124p. Ridgway, K.R. (2007) Long-term trend and decadal variability of the southward penetration of the East Australian Current. Geophysical Research Letters 34: 22921-22936. Robinson, A. C., and Dennis, T. E. (1988) The status and management f seal populations in South Australia. In Augee, M. L. (Ed.) Marine Mammals of Australasia – Field Biology and Captive Management. pp. 87–104. Royal Zoological Society of New South Wales. Rowling, K. (2007) Caulerpa taxifolia - 2007 survey of current distribution and high risk areas. Prepared for PIRSA Biosecurity. SARDI Research Report Series No: 234. Russell, B.C. (1986) A new species of Suezichthys (Pisces: Labridae) from the Great Australian Bight. Transactions of the Royal Society of South Australia 110(1-2): 59-61. Russell-French, A. (2002) Question on Notice-Review of Australia's Quarantine Function. Letter to Ms Margot Kerley, Secretary, Joint Committee of Public Accounts and Audit. Department of the Environment and Heritage, Canberra. Parsons, M.J.G., McCauley, R.D. and Mackie, M.. (2007) Characterisation of a mulloway (Argyrosomus japonicus) spawning aggregation using passive acoustic techniques. Presented at the 2nd International Conference and Exhibition on Underwater Acoustic Measurements: Technologies and Results, Crete, June, 2007. Parsons, M. J., McCauley, R. D., Mackie, M. C., Siwabessy, P., and Duncan, A. J. (2009) Localization of individual mulloway (Argyrosomus japonicus) within a spawning aggregation and their behaviour throughout a diel spawning period. ICES Journal of Marine Science 66: 1007–1014. Potter, A., Southby, C. and Heap, A.D., (2006) Geomorphology and Sedimentology of the South West Planning Region of Australia. Geoscience Australia report. Geoscience Australia, Canberra. Robinson, A.C. and Dennis, T.E. (1988). The status and management of seal populations in South Australia In: Augee, M.L (Ed.) Marine Mammals of Australasia: field biology and captive management. Royal Zoological Society of NSW, Sydney. Rodda, K.R. (2000) Development in the Port Jackson shark embryo. Ph.D. thesis, University of Adelaide, South Australia. 238p. Rosenzweig, C., Karoly, D., Vicarelli, M., Neofotis, P., Wu, Q., Casassa, G., Menzel, A., Root, T.L., Estrella, N., Seguin, B., Tryjanowski, P., Liu, C., Rawlins, S., and Imeson, A. (2008) Attributing physical and biological impacts to anthropogenic climate change. Nature 453: 353-357. Russell, B.D., Thompson, J.I., Falkenberg, L.J. and Connell, S.D. (2009) Synergistic effects of climate change and local stressors: CO2 and nutrient driven change in subtidal rocky habitats. Global Change Biology 15: 2153–2162. SARDI Aquatic Sciences (2006) Tagging sea lions hits new heights at the Bight. SARDI Communicator, Autumn 2006. Scheibling, R.E. and Gagnon. P. (2009) Temperature-mediated outbreak dynamics of the invasive bryozoan Membranipora membranacea in Nova Scotian kelp beds. Marine Ecology Progress Series 390: 1–13. 132 Seddon, S. Connolly, R and Edyvane, K.S. (2000) Large-scale seagrass dieback in northern Spencer Gulf, South Australia. Aquatic Botany 66(4): 297-310. Sfriso, A. (2010). Coexistence of Ulva rigida and Ulva laetevirens (Ulvales, Chlorophyta) in Venice Lagoon and other Italian transitional and marine environments. Botanica Marina 53: 9-18. Shanks, S. (2004) Ecological Assessment of the South Australian Pilchard Fishery. Assessment report prepared for Department of Environment and Heritage, against the Guidelines for the Ecologically Sustainable Management Of Fisheries for the purposes of Part 13 and 13(A) of the Environment Protection and Biodiversity Conservation Act 1999. Primary Industries & Resources South Australia in association with the pilchard industry and the South Australian Research and Development Institute. Shaughnessy, P. and Dennis, T. (2002). Population assessment of some colonies of New Zealand fur seals and Australian sea lions in South Australia, 2001–2002. Report to National Parks and Wildlife South Australia, Department for Environment and Heritage. 28 p. Shaughnessy, P., Gales, N., Dennis, T. and Goldsworthy, S. (1994) Distribution and abundance of New Zealand fur-seals, Arctocephalus forsteri, in South Australia and Western Australia. Wildlife Research 21: 667-95. Shaughnessy, P., Dennis, T. and Saeger, P. (1997) Abundance, seasonality of breeding and rate of entanglement of Australian sea lions (Neophoca cinerea) at colonies on the west coast of South Australia. Report to Biodiversity Group, Environment Australia, Canberra. 18p. Shaughnessy, P.D., Dennis, T.E. and Saeger, P.G. (2005) Status of Australian sea lions, Neophoca cinerea, and New Zealand fur seals Arctocephalus fosteri, on Eyre Peninsula and the Far West Coast of South Australia. Wildlife Research 32: 85-101. Shaughnessy, P.D., Kirkwood, R., Cawthorn, M., Kemper, C. and Pemberton, D. (2003) Pinnipeds, cetaceans, and fisheries in Australia: A review of operational procedures. In: Gales, N., Hindell, M. and Kirkwood, R. (Eds) Marine Mammals: Fisheries, tourism and management issues. Melbourne, CSIRO: 136-152. Shephard, J., Catterall, C. and Hughes, J. (2005) Long-term variation in the distribution of the Whitebellied Sea-eagle (Haliaeetus leucogaster) across Australia. Austral Ecology 30: 131-145. Shepherd, S.A. and Baker, J.L. (2008) Reef fishes of lower Gulf St Vincent. Chapter in: Shepherd, S.A., Bryars, S., Kirkegaard, I.R., Harbison, P. and Jennings, J. (2008) The Natural History of Gulf St Vincent. Royal Society of South Australia. Short, A.D., Fotheringham, D. and Buckley, R.C. (1986) Coastal Morphodynamics and Holocene Evolution of the Eyre Peninsula Coast, South Australia. Coastal Studies Unit Technical Report No. 86/ 2, December 1986. University of Sydney, New South Wales. Simpfendorfer, C.A. and McAuley, R. (2003) Furgaleus macki. In: IUCN (2009) IUCN Red List of Threatened Species. Version 2009.2. www.redlist.org. Simpfendorfer, C.A. and Unsworth, P. (1998) Gillnet mesh selectivity of dusky (Carcharhinus obscurus) and whiskery (Furgaleus macki) sharks from south-western Australia. Marine and Freshwater Research 49: 713-18. Simpfendorfer, C.A. and Unsworth, P. (1998) Reproductive biology of the whiskery shark, Furgaleus macki, off south-western Australia. Marine and Freshwater Research 49: 687-793. 133 Simpfendorfer, C.A., Goodreid, A., and McAuley, R.B. (2001) Diet of three commercially important shark species from Western Australian waters. Marine and Freshwater Research 52: 975 – 985. Simpfendorfer, C.A., McAuley, R., Chidlow, J.A., Lenanton, R., Hall, N.G. and Bastow, T. (1999) Biology and Stock Assessment of Western Australia’s Commercially Important Shark Species. Final Report to the Fisheries Research and Development Corporation. Smale, M.J. (1991) Occurrence and feeding of three shark species, Carcharhinus brachyurus, C. obscurus and Sphyrna zygaena, on the eastern Cape Coast of South Africa. South African Journal of Marine Science 11: 31-42. Sorokin, S.J., Currie, D. and Ward, T.M. (2005) Sponges from the Great Australian Bight Marine Park Benthic Protection Zone. SARDI Aquatic Sciences Publication No. RD04/1070 for Wildlife Conservation Fund, South Australian National Parks and Wildlife Council. Sorokin, S.J., Fromont, J. and Currie, D. (2007) Demosponge biodiversity in the Benthic Protection Zone of the Great Australian Bight. Transactions of the Royal Society of South Australia 131(2): 192-204. Sorokin,S.J., Laperousaz, T.C.D. and Drabsch, S.L. (2007) A Catalogue of Shallow-water Sponges from the Investigator Islands, South Australia. SARDI Research Report Series No: 258. SARDI Aquatic Sciences, South Australia. South Australian Government (2004) Natural Resources Management Act 2004. http://www.austlii.edu.au/au/legis/sa/consol_act/nrma2004298/ Stafford, H. and Willan, R.C. (2007) Is it a Pest? Introduced and naturalised marine animal species of Torres Strait Northern Australia. Queensland Department of Primary Industries and Fisheries, Cairns, Queensland. Stevens, J.D. (1984) Biological observations on sharks caught by sport fishermen off New South Wales. Australian Journal of Marine and Freshwater Research 35: 573-590. Stevens, J.D. (1999) Variable resilience to fishing pressure in two sharks: the significance of different ecological and life history parameters. American Fisheries Society Symposium 23: 11-15. Stevens, J.D. and West, G. (1997) Investigation of School and Gummy Shark Nursery Areas in Southeastern Australia. FRDC Report Project No. 93/061. 76p. Stuart-Smith, R.D., Barrett, N.S., Stevenson, D.G. and Edgar, G.J. (2009) Stability in temperate reef communities over a decadal time scale despite concurrent ocean warming. Global Change Biology doi: 10.1111/j.1365-2486.2009.01955.x Struckmeyer, H.I.M. (2009) Ceduna Sub-basin, Bight Basin: Results of 3D Petroleum Systems Modelling. Ceduna 3D PS Model –September 2009 (Geocat No. 69485). Geoscience Australia, Canberra. Svane, I., Rodda, K. and Thomas, P. (2007) Prawn fishery by-catch and discards: marine ecosystem analysis – population effects. SARDI Aquatic Sciences Publication No. RD 03-0132. Report No. 199. 404p. Talman, S.G., Brown, L.P. and Gason, A.S.H. (2005) Integrated Scientific Monitoring Program: Great Australian Bight Trawl Fishery Annual Report 2004. PIRVic (Primary Industries Research Victoria) report for Australian Fisheries Management Authority. 134 The Fishing Guide (2009) Surf Fishing for Mulloway at Yalata. Article by Pete. http://www.thefishingguide.com.au/index.php/fishing-reports/south-australia/286-surf-fishing-formulloway-at-yalata.html (URL accessed December 2009) Treloar, M.A. (2006) Urolophus expansus. In: IUCN (2009) IUCN Red List of Threatened Species. Version 2009.2. www.iucnredlist.org. Trinder, D. (2007) Chapter 4.18 Demersal fish – slope. In: McClatchie, S. et al. (Eds) Review of Ecological Information and Knowledge of Australia’s South West Marine Region (SWMR). Report by SARDI Aquatic Sciences and University of Western Australia, for National Oceans Office, Canberra. http://www.environment.gov.au/coasts/mbp/publications/south-west/pubs/sw-ecosystems-part2.pdf Verco, J.C. (1912) Shells from the Great Australian Bight. Transactions of the Royal Society of South Australia 36: 206–232. Vuong, V., Capon, R.J., Lacey, E., Gill, J.H., Heiland, K. and Friedel, T. (2001) Onnamide F: A new nematocide from a southern Australian marine sponge Trachycladus laevispirulifer. J. Nat. Prod. 64 (5): 640–642. Walker, T.I. (1996) Stock Assessment Report. Gummy shark 1995. Compiled for the Southern Shark Fishery Assessment Group. Australian Fisheries Management Authority, Canberra. Walker, T.I. (2007) Spatial and temporal variation in the reproductive biology of gummy shark Mustelus antarcticus (Chondrichthyes : Triakidae) harvested off southern Australia. Marine and Freshwater Research 58: 67-97. Walker, T.I., Moulton, P., Dow, N. and Saddlier, S. (1989) Southern Shark Assessment Project. Fishing Industry Research Trust Account Final Report. Internal Report No. 175. Marine Science Laboratories, Victorian Fisheries Research Institute, Queenscliff. 43p. Walker, T.I., Taylor, B. and Hudson, R. (1999) Southern Shark Catch and Effort 1970-98. Report to Australian Fisheries Management Authority. Marine and Freshwater Resources Institute, Queenscliff, Victoria. 34p. Walker, T.I., Punt, A., Taylor, B.L. and Brown, L.P. (2000) Modelling school shark (Galeorhinus galeus) movement in the Southern Shark Fishery. In: Hancock, D.A., Smith, D.C. and Koehn, J.D. (Eds) Fish Movement and Migration. Proceedings of the Australian Society for Fish Biology conference, 28-29th September 1999, Bendigo, Victoria. Australian Society for Fish Biology, Sydney, New South Wales. Walker, T.I., Taylor, B. and Brown, L. (2000) Southern Shark Tag Database Project, FRDC Project No. 96/162, Draft Final Report to Fisheries Research and Development Corporation. Marine and Freshwater Resources Institute, Queenscliff, Victoria. 79p. Walter, J.P. and Ebert, D.A. (1991) Preliminary estimates of age of the bronze whaler Carcharhinus brachyurus (Chondrichthyes: Carcharhinidae) from southern Africa, with a review of some life history parameters. South African Journal of Marine Science 10: 37-44. Walters, L.J. and Abgrall, M.J. (2000) Impact of drift species (Gracilaria compressa, Zoobotryon verticillatum) on the recruitment and fitness of sessile invertebrates. American Zoologist 40(6): 12491250. Ward, T.J. and Butler, A. (2006) Coasts and oceans. Theme commentary prepared for the 2006 Australia State of the Environment Committee, Department of Environment and Heritage, Canberra. http://www.deh.gov.au/soe/2006/commentaries/coasts/index.html 135 Ward, T.M., McLeay, L. and Rogers, P. (2003) Benthic Protected Zone of the Great Australian Bight Marine Park. 2: Monitoring Sustainable Use. South Australian Research and Development Institute report, for National Parks and Wildlife South Australia, and the Commonwealth Department of the Environment and Heritage. SARDI Aquatic Sciences, South Australia. http://www.deh.gov.au/coasts/mpa/gab/monitor/pubs/monitor.pdf Ward, T.M., McLeay, L. and Rogers, P. (2005) Spawning biomass of sardine (pilchard, Sardinops sagax) in South Australian waters in 2005. Report to PIRSA Fisheries. SARDI Aquatic Sciences Publication No. RD05/0019-1, Research Report Series No. 101. SARDI Aquatic Sciences, South Australia. Ward, T.M., Kinloch, M., Jones G.K., and Neira, F.J. (1998) A collaborative investigation of the usage and stock assessment of baitfish in southern and eastern Australian waters, with special reference to pilchards (Sardinops sagax). Final Report to FRDC. 324p. Ward, T.M., Hoedt, F., McLeay, L., Dimmlich, W.F., Jackson, G., Rogers, P.J. and Jones, G.K. (2001) Have recent mass mortalities of the sardine Sardinops sagax facilitated an expansion in the distribution and abundance of the anchovy Engraulis australis in South Australia? Marine Ecology Progress Series 220: 241–251. Ward, T.M., Goldsworthy, S.D., Rogers, P.J., Page, B., McLeay, L.J., Dimmich, W., Baylis, A.M.M., Einoder, L., Wiebkin, A., Roberts, M., Daly, K., Caines, R. and Huveneers, C. (2008) Ecological importance of small pelagic fishes in the Flinders Current system. SARDI Aquatic Sciences report to the Commonwealth Department of the Environment, Water, Heritage and the Arts. SARDI Aquatic Sciences Publication F2007/001194-1. Ward, T.M., McLeay, L.J., Dimmlich, W.F., Rogers, P.J., McClatchie, S., Matthews, R., Kaempf, J. and van Ruth, P.D. (2006) Pelagic ecology of a northern boundary current system: effects of upwelling on the production and distribution of sardine (Sardinops sagax), anchovy (Engraulis australis) and southern bluefin tuna (Thunnus maccoyii) in the Great Australian Bight. Fisheries Oceanography 15(3): 191–207. Ward, T.M., Sorokin, S.J., Currie, D.R., Rogers, P.J. and McLeay, L.J. (2006) Epifaunal assemblages of the eastern Great Australian Bight: Effectiveness of a benthic protection zone in representing regional biodiversity. Continental Shelf Research 26: 25–40. Ward, T.M., Sorokin, S.J., Rogers, P.J., McLeay, L.J. and Turner, D.J. (2003) Benthic Protected Zone of the Great Australian Bight Marine Park: 3. Pilot Study for Performance Assessment (Volume 1). Final Report to National Parks and Wildlife South Australia and the Commonwealth Department for Environment and Heritage. SARDI Aquatic Sciences, South Australia. Ward, T.M., Sorokin, S.J., Rogers, P.J., McLeay, L.J. and Turner, D.J. (2003) Benthic Protected Zone of the Great Australian Bight Marine Park: 3. Pilot Study for Performance Assessment (Volume 2). Final Report to National Parks and Wildlife South Australia and the Commonwealth Department for Environment and Heritage. SARDI Aquatic Sciences, South Australia. Watters, G. (2004). Digital Murex: Viator. Division of Molluscs, Museum of Biological Diversity, Department of Evolution, Ecology, and Organismal Biology, Ohio State University, Ohio. Webb, J.A., and James, J.M. (2006) Karst evolution of the Nullarbor Plain, Australia, in Harmon, R.S., and Wicks, C. (Eds) Perspectives on Karst Geomorphology, Hydrology, and Geochemistry. A Tribute Volume To Derek C. Ford and William B. White. Geological Society of America Special Paper 404, p. 65– 78. Wernberg, T., Campbell, A., Coleman, M.A., Connell, S.D., Kendrick, G.A., Moore, P.J., Russell, B.D., Smale, D. and Steinberg, P.D. (2009) Macroalgae and temperate rocky reefs. In: Poloczanska, E.S., Hobday, A.J. and Richardson, A.J. (Eds) A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009. NCCARF Publication 05/09, ISBN 978-1-921609-03-9. 136 Wilson, B., Wilson, C. and Baker, P. (1993). Australian Marine Shells: Prosobranch Gastropods Part 1. Odyssey Press, Western Australia. Wilson, B., Wilson, C. and Baker, P. (1994). Australian Marine Shells: Prosobranch Gastropods Part 2. Odyssey Press, Western Australia. Wilson, D., Curtotti, R., Begg, G. and Phillips, K. (Eds) (2009) Fishery Status Reports 2008: Status Of Fish Stocks And Fisheries Managed By The Australian Government. Bureau of Rural Sciences & Australian Bureau of Agricultural and Resource Economics, Canberra. Wilson, G. (2007) Yalata: to be or not to be. Fishing article in fishnet web site: http://www.fishnet.com.au/default.aspx?id=234&articleId=6122 (URL accessed December 2009). Wiltshire, K. and Rowling, K. (2009). Caulerpa taxifolia - 2009 surveys of current distribution and high risk areas. SARDI Publication No. F2009/000347-1. SARDI Research Report Series No. 369. Womersley, H.B.S. and Edmunds, S.J. (1958) A general account of the intertidal ecology of South Australian coasts. Australian Journal of Marine and Freshwater Research 9(2): 217-260. Yalata Land Management (2003) The Community: http://www.yalata.org/default.htm The Lands: http://www.yalata.org/lands.htm The Whales: http://www.yalata.org/whales.htm Camping and Fishing: http://www.yalata.org/camping.htm Permits and Bookings: http://www.yalata.org/permits.htm Young, J.W., Lamb, T.D., Le, D., Bradford, R.W. and Whitelaw, A.W. (1997) Feeding ecology and inter-annual variations in diet of southern bluefin tuna, Thunnus maccoyii, in relation to coastal and oceanic waters off eastern Tasmania, Australia. Env. Biol. Fishes 50: 275-291. Young, J.W. , Nishida, S,. and Stanley, C. (2000) A preliminary survey of the summer hydrography and plankton biomass of the eastern Great Australian Bight, Australia. Southern Bluefin Tuna Recruitment Monitoring and Tagging Program: Report of the Eleventh Workshop, CRIMP, Hobart, Tas. (Australia). 137 7. Appendices Appendix 1: Decapod crustacean species known from the Great Australian Bight upper and mid shelf waters, and thus highly likely to occur in the AW NRM marine region (Davie, 2002, in ABRS, 2009; Ward et al., 2003; Poore, 2004; Currie et al., 2007; South Australian Museum, Museum Victoria and W.A. Museum records, cited in OZCAM database, 2009). Common names collectively as listed in Davie (2002, in ABRS, 2009) and CSIRO (2009). E = apparently endemic within South Australia. A = recorded directly east, west and/or south of AW NRM, and therefore likely to occur in the region. U = might occur in AW NRM region, at upper depth limit. An additional species Leptomithrax globifer Rathbun, 1918 that occurs in GAB waters is known from 150m – 220m (Davie, in ABRS, 2009) and is excluded from the list below, because its presence in AW NRM region is unknown. Scientific Name Common Name Actaea calculosa (H. Milne Edwards, 1834) Actaea peronii H. Milne Edwards, 1834 / Actaea peronii peronii (H. Milne Edwards, 1834) Acutigebia simsoni (Thomson, 1893) Alope orientalis (de Man, 1890) Alpheopsis trispinosa (Stimpson, 1861) Alpheus bidens (Olivier, 1811) A Alpheus edwardsii (Audouin, 1827) Alpheus novaezealandiae Miers, 1876 A Alpheus papillosus Banner & Banner, 1982 Alpheus parasocialis Banner & Banner, 1982 Alpheus villosus (Olivier, 1811) Amarinus laevis (Targioni Tozzetti, 1877) A Anchistus custos (Forskål, 1775) Arcania undecimspinosa De Haan, 1841 A Aristaeomorpha foliacea (Risso, 1827) Athanas granti Coutière, 1908 Athanas haswelli / hasswelli Coutière, 1908 Austrodromidia australis (Rathbun, 1923) U Austrodromidia insignis (Rathbun, 1923) A, U Austrodromidia octodentata (Haswell, 1882) Bellidilia laevis (Bell, 1855) A Bellidilia undecimspinosa Kinahan, 1856 Betaeus australis Stimpson, 1860 Biffarius ceramicus (Fulton & Grant, 1906) A Calliax aequimana (Baker, 1907) A Cancellus typus H. Milne Edwards, 1836 A Carcinoplax meridionalis Rathbun, 1923 = Pycnoplax meridionalis (Rathbun, 1923) U Ceratocarcinus longimanus White, 1847 Ceratoplax glaberrima (Haswell, 1881) Cyclograpsus audouinii H. Milne Edwards, 1837 Cyclograpsus granulosus H. Milne Edwards, 1853 Chlorotocella spinicaudus (H. Milne Edwards, 1837) Cyclograpsus granulosus H. Milne Edwards, 1853 Dagnaudus petterdi (Grant, 1905) U Dardanus arrosor (Herbst, 1796) Dorhynchus ramusculus (Baker, 1906) Dromidiopsis globosa (Lamarck, 1818) Elamena abrolhensis Gordon, 1940 Facetted Crab Spiky Stone Crab / Thorn-legged crab Sand-borer Bald Prawn (a prawn) Mottled Snapping Prawn (a pistol prawn / snapping prawn) New Zealand Snapping Prawn (a pistol prawn / snapping prawn) (a pistol prawn / snapping prawn) Hairy Pistol Prawn (a flat spider crab) Pinna Prawn (a pebble crab) Giant Red Prawn, Red Prawn, Royal Red Prawn (a snapping shrimp) (a snapping shrimp) Southern Sponge Crab Adorned Sponge Crab Bristled Sponge-crab Smooth Pebble Crab Large Pebble Crab (a prawn / shrimp) Ceramic Ghost Shrimp, Large-clawed Burrower Ambidextrous Burrower Miner Hermit crab (a goneplacid crab) Gouty Crab (a hairy crab) (a shore crab) (a shore crab) Slender-beaked Prawn Rough Shore Crab Antlered Crab Striated Hermit crab Slender Spider Crab Shaggy Sponge Crab (a flat spider crab) 138 Ebalia crassipes (Bell, 1855) A Ebalia dentifrons Miers, 1886 A Ebalia intermedia Miers, 1886 A Ebalia tuberculosa (A. Milne Edwards, 1873) Ephippias endeavouri Rathbun, 1918 Fultodromia nodipes (Guérin-Méneville, 1832) Galathea australiensis Stimpson, 1858 Galathea magnifica Haswell, 1882 Georgeoplax glabra (Baker, 1906) A Haledromia bicavernosa (Zietz, 1888) E Halicarcinides nuytsi (Hale, 1927) A Halicarcinus ovatus Stimpson, 1858 Halicarcinus rostratus (Haswell, 1881) Hippa australis Hale, 1927 Hippolyte australiensis Stimpson, 1860 A Hippolyte caradina Holthuis, 1947 Huenia australis Griffin & Tranter, 1986 A Huenia halei Griffin & Tranter, 1986 Hypothalassia acerba Koh & Ng, 2000 Ibacus alticrenatus Bate, 1888 Ibacus peronii Leach, 1815 Leptograpsodes octodentatus (H. Milne Edwards, 1837) Leptomithrax species (in Ward et al., 2003) Jasus (Jasus) edwardsii (Hutton, 1875) Latreutes compressus (Stimpson, 1860) Leander tenuicornis (Say, 1818) Leptochela (Leptochela) sydniensis Dakin & Colefax, 1940 Leptomithrax sternocostulatus (H. Milne Edwards, 1851) Leptograpsodes octodentatus (H. Milne Edwards, 1837) Leptograpsus variegatus (Fabricius, 1793) Leptomithrax gaimardii (H. Milne Edwards, 1834) Leptomithrax sternocostulatus (H. Milne Edwards, 1851) Lomis hirta (Lamarck, 1818) Macrobrachium intermedium (Stimpson, 1860) Megametope carinata (Baker, 1907) A Melicertus latisulcatus (Kishinouye, 1896) = Penaeus (Melicertus) latisulcatus Kishinouye, 1896 Merocryptus lambriformis A. Milne Edwards, 1873 Metanephrops boschmai (Holthuis, 1964) U Metanephrops velutinus Chan & Yu, 1991 U Metapenaeopsis lindae Manning, 1988 A Micippa philyra (Herbst, 1803) A Micippa thalia (Herbst, 1803) A Micropagurus acantholepis (Stimpson, 1858) Munida haswelli Henderson, 1885 Naxia aurita (Latreille, 1825) Naxia spinosa (Hess, 1865) Nectocarcinus integrifrons (Latreille, 1825) Nectocarcinus spinifrons Stephenson, 1961 Nectocarcinus tuberculosus A. Milne Edwards, 1860 A Neocallichirus angelikae Sakai, 2000 E, A Notopontonia platycheles Bruce, 1991 Ogyrides delli Yaldwyn, 1971 (a nut crab / pebble crab) (a nut crab / pebble crab) Smooth Nut Crab Nut Crab / Pebble Crab Endeavour Crab Warty Sponge Crab Striated Craylet Scaled Craylet (a crab) Eared Sponge Crab Beakless Sea-spider Three-pronged Sea-spider Beaked Sea-spider / Spider Crab Mole Crab / Sand Crab Southern Weed Shrimp (a shrimp) (a spider crab) (a spider crab) Champagne Crab Deep Water Bug Balmain Bug Burrowing Shore Crab (a spider crab) Southern Rock Lobster Green Prawn (a carid shrimp) Sydney Comb Shrimp Ribbed Spider-crab Burrowing Shore Crab (a shore crab) Great Spider Crab Ribbed Spider Crab Hairy Stone Crab Striped Prawn Crested Forehead Crab Western King Prawn (a pebble crab) Bight Lobster / Bight Scampi / Boschma's Scampi Andaman Scampi / Velvet Lobster / Velvet Scampi Linda‘s Velvet Prawn (a flat-beaked crab) (a flat-beaked crab) (a hermit crab) Long-armed Craylet Smooth Seaweed Crab Spiny Seaweed Crab Rough Rock Crab (a rock crab) Rough Rock Crab (a slow prawn) (a carid shrimp) (a shrimp) (NB known from E and SE Australia, but Currie et al. 2007 reported this species during a survey of the GAB Marine Park) Oplophorus novaezeelandiae (de Man, 1931) (a carid shimp) 139 Ovalipes australiensis Stephenson & Rees, 1968 Ozius truncatus H. Milne Edwards, 1834 A Paguristes brevirostris Baker, 1905 Paguristes frontalis (H. Milne Edwards, 1836) A Paguristes tuberculatus Whitelegge, 1900 A Paguristes sulcatus Baker, 1905 A Pagurixus handrecki Gunn & Morgan, 1992 A Pagurixus jerviensis McLaughlin & Haig, 1984 A Pagurus sinuatus (Stimpson, 1858) A Palaemon litoreus (McCulloch, 1909) A Palaemon serenus (Heller, 1862) Paragrapsus gaimardii (H. Milne Edwards, 1837) Paramithrax barbicornis (Latreille, 1825) Periclimenes aesopius (Bate, 1863) Petrocheles australiensis (Miers, 1876) Philocheras intermedius (Bate, 1863) Phylladiorhynchus pusillus (Henderson, 1885) Pilumnus acer Rathbun, 1923 Pilumnus fissifrons Stimpson, 1858 A Pilumnus rufopunctatus Stimpson, 1858 A Pilumnus tomentosus Latreille, 1825 Pinnotheres hickmani (Guiler, 1950) Pisidia dispar (Stimpson, 1858) Plagusia chabrus (Linnaeus, 1758) A Polyonyx transversus (Haswell, 1882) Portunus (Portunus) pelagicus (Linnaeus, 1758) Portunus (Portunus) sanguinolentus sanguinolentus (Herbst, 1783) A Prismatopus spatulifer (Haswell, 1881) Processa gracilis Baker, 1907 A Pseudocarcinus gigas (Lamarck, 1818) Pseudopalicus macromeles Castro 2000 Pseudopontonia minuta (Baker, 1907) Rhynchocinetes enigma Okuno, 1997 Rochinia mosaica (Whitelegge, 1900) Schizophrys dama (Herbst, 1804) Schizophrys rufescens Griffin & Tranter, 1986 Scyllarus mawsoni (Bage, 1938) Scyllarus / Scyllarides species 1 (in Ward et al., 2003) Sergestes sargassi Ortmann, 1893 Stimdromia lamellata (Ortmann, 1894) Stimdromia lateralis (Gray, 1831) Strahlaxius waroona (Poore & Griffin, 1979) A Strigopagurus elongatus Forest, 1995 A Synalpheus fossor (Paul'son, 1875) A Synalpheus harpagatrus Banner & Banner, 1975 Synalpheus streptodactylus Coutière, 1905 Synalpheus tumidomanus (Paul'son, 1875) Thalamita intermedia Miers, 1886 Trigonoplax longirostris McCulloch, 1908 Thalamita sima H. Milne Edwards, 1834 A Upogebia bowerbankii (Miers, 1884) A Upogebia tractabilis (Hale, 1941) A Vercoia gibbosa Baker, 1904 Xanthid crab 1 (in Ward et al., 2003) Sand Crab Reef Crab Southern Hermit Crab Common Hermit Crab Friendly Hermit Crab Hairy-legged Hermit Crab Clarrie‘s Hermit Crab (a hermit crab) (a hermit crab) Shore Prawn Rock-pool Prawn Common Shore Crab Sea Toad Aesop Prawn / Cleaner Shrimp Spiny Porcelain Crab (a carid shrimp) Little Craylet Long-spined Hairy Crab Tasselled Crab Red-spotted Hairy Crab Common Hairy Crab Pea Crab Little Porcelain Crab Cleft-fronted Shore Crab Polished Porcelain Crab Blue Swimmer Crab / Blue Swimming Crab Red-spotted Swimming Crab (spider crab) Long-wristed Shrimp Queen Crab / Tasmanian Giant Crab (a crab) Smooth Prawn (a prawn / shrimp) Little Thornback Crab (a decorator crab) (a spider crab) (a lobster) (a slipper lobster) (a penaeid prawn) (a sponge crab) Ridged Sponge Crab (a slow prawn) (a hermit crab) (a snapping shrimp) (a snapping shrimp) (a snapping shrimp) Fat-handed Shrimp Six-lobed Crab (a flat spider crab) Four-lobed Swimming Crab (a slow prawn) (a slow prawn) Humped Shrimp (a crab) 140 Appendix 2: Echinoderm species known from the Great Australian Bight upper and mid shelf waters, and thus highly likely to occur in the AW NRM marine region (Rowe and Gates, 1995, and O‘Hara, 2001, in ABRS, 2009; O‘Loughlin 2002; O‘Loughlin et al., 2003; Ward et al., 2003; O‘Loughlin and Waters, 2004). Common names as listed in Edgar (2008) and/or ABRS (2009). An additional three species Amphiophiura urbana (Koehler, 1904), Ophiocreas sibogae Koehler 1904, and Pseudostichopus mollis Théel 1886, all of which occur on the continental slope, have reported upper depth limits of about 90100m, and it is not known whether they also occur in the waters of the AW NRM region. Another species, Ophiomusium anisacanthum H.L. Clark 1928, widespread across southern Australia, is known from 130m – 310m and thus not included in the table below, nor are Ophiomusium australe H.L. Clark 1928, known to date from 130m – 595m, and Ophiozonella bispinosa (Koehler, 1897), known from 205m – 1,277m. Scientific Name Common Name Allostichaster polyplax (Müller & Troschel, 1844) Amblypneustes elevatus (Hutton, 1872) Amblypneustes formosus Valenciennes, 1846 Amblypneustes grandis H.L. Clark, 1912 Amblypneustes pallidus (Lamarck, 1816) Amphioplus (Amphichilus) ochroleuca (Brock, 1888) Amphipholis squamata (Delle-Chiaje, 1828) Amphistigma minuta H.L. Clark, 1938 Amphiura (Amphiura) constricta Lyman, 1879 Amphiura (Amphiura) multiremula H.L. Clark, 1938 Amphiura (Ophiopeltis) parviscutata A.M. Clark, 1966 Antedon incommoda Bell, 1888 Anthaster valvulatus (Müller & Troschel, 1843) Aporometra occidentalis H.L. Clark, 1938 Asterodiscides truncatus (Coleman, 1911) Astropecten preissi Müller & Troschel, 1843 Astroboa ernae Döderlein, 1911 Astropecten vappa Müller & Troschel, 1843 Astrosierra microconus (H.L. Clark, 1914) Australocnus occiduus (O'Loughlin & O'Hara, 1992) Bollonaster pectinatus (Sladen, 1883) Brissus (Allobrissus) agassizii Döderlein, 1885 Cenolia spanoschistum H.L. Clark, 1916 Cenolia tasmaniae (A.H. Clark, 1918) = Comanthus tasmaniae (A.H. Clark, 1918) Cenolia trichoptera (Müller, 1846) = Comanthus trichoptera Centrostephanus tenuispinus H.L. Clark, 1914 Centrostephanus sp. (in Ward et al., 2003) Cercodemas anceps Selenka, 1867 Ceto cuvieria (Cuvier, 1817) Clarkcoma canaliculata (Lütken, 1869) Clarkcoma pulchra (H.L. Clark, 1928) Colochirus crassus Ekman, 1918 Colochirus quadrangularis Troschel, 1846 Comatulella brachiolata (Lamarck, 1816) Conocladus australis (Verrill, 1876) Coscinasterias muricata Verrill, 1867 Cucuvitrum rowei O'Loughlin & O'Hara, 1992 Dipsacaster magnificus (H.L. Clark, 1916) Echinaster arcystatus H.L. Clark, 1914 Echinaster glomeratus H.L. Clark, 1916 Echinocardium cordatum (Pennant, 1777) 141 Many-armed Seastar (a sea urchin) (a sea urchin) (a sea urchin) Yellow-spined Egg Urchin (a brittle star) (a brittle star) (a brittle star) (a brittle star) (a brittle star) (a brittle star) Variable Feather Star Mottled Seastar (a feather star / crinoid) Firebrick Seastar Preiss‘s Sand Star Pink Basket Star Comb Sand Star (a basket star) (a sea cucumber) (a sea star) Agassiz‘s Heart Urchin (a feather star) Tasmanian Feather Star Orange Feather Star Western Hollow-spined Urchin (a sea urchin) Red Box Sea Cucumber (a black and white sea cucumber) Variable Brittle Star White-flecked Brittle Star (a sea cucumber) Thorny Sea Cucumber (a feather star on macroalgae & in seagrass) Southern Basket Star Eleven-armed Seastar (a sea cucumber) (a sea star) Pale Mosaic Seastar Orange Seastar Heart Urchin Echinocyamus platytatus H.L. Clark, 1914 Euryale asperum Lamarck, 1816 Fibularia (Fibularia) plateia H.L. Clark, 1928 Fromia polypora H.L. Clark, 1916 Goniocidaris tubaria (Lamarck, 1816) (green mottled aspidochirotida, in Ward et al., 2003) Heliocidaris erythrogramma (Valenciennes, 1846) Henricia compacta (Sladen, 1889) Holopneustes inflatus (Lütken, 1872) Holopneustes porosissimus L. Agassiz, 1846 Holopneustes purpurascens A. Agassiz, 1872 Holothuria (Thymiosycia) hartmeyeri Erwe, 1913 Leptosynapta dolabrifera (Stimpson, 1855) Lipotrapeza vestiens (Joshua, 1914) Luidia australiae Döderlein, 1920 Mensamaria intercedens (Lampert, 1885) Meridiastra atyphoida ( H.L. Clark, 1916) = Asterina atyphoida H.L. Clark, 1916 Meridiastra calcar (Lamarck, 1816) = Patiriella calcar (Lamarck, 1816) Meridiastra gunnii (Gray, 1840) = Patiriella gunnii (Gray, 1840) = Patiriella brevispina H.L. Clark, 1938 Meridiastra medius O‘Loughlin, Waters and Roy, 2003 Meridiastra occidens O‘Loughlin, Waters and Roy, 2003 Meridiastra oriens O‘Loughlin, Waters and Roy, 2003 Moira lethe Mortensen, 1930 Nectria macrobrachia H.L. Clark, 1923 Nectria multispina H.L. Clark, 1928 Nectria ocellata Perrier, 1876 Nectria pedicelligera Mortensen, 1925 Nectria saoria Shepherd, 1967 Nectria wilsoni Shepherd & Hodgkin, 1965 Neoamphicyclus lividus Hickman, 1962 Neoamphicylcus mutans (Joshua, 1914) = Cucumella mutans (Joshua, 1914) Nepanthia troughtoni (Livingstone, 1934) = Pseudonepanthia troughtoni (Livingstone, 1934) Odontohenricia endeavouri Rowe & Albertson, 1988 Ophiacantha alternata A.M. Clark, 1966 Ophiacantha clavigera Koehler, 1907 Ophiactis resiliens Lyman, 1879 Ophiactis tricolor H.L. Clark, 1928 Ophiarachnella ramsayi (Bell, 1888) Ophiocrossota multispina (Ljungman, 1867) Ophioconus opacum (H.L. Clark, 1928) Ophiomyxa australis Lütken, 1869 Ophionereis schayeri (Müller & Troschel, 1844) Ophionereis semoni (Döderlein, 1896) Ophiopeza cylindrica (Hutton, 1872) Ophiopsammus assimilis (Bell, 1888) Ophiothrix (Keystonea) hymenacantha H.L. Clark, 1928 E (known only from type locality) Ophiothrix (Ophiothrix) caespitosa Lyman, 1879 Ophiothrix (Placophiothrix) albostriata H.L. Clark, 1928 E (known only from type locality) Ophiothrix (Placophiothrix) spongicola Stimpson, 1855 142 (a sand dollar) (a snake star) / Spiny Basket Star (a sand dollar) Many-spotted Seastar Stumpy Pencil Urchin (a sea cucumber) Purple Urchin (a sea star) Inflated Egg Urchin Carmine-spotted Egg Urchin Eastern Egg Urchin (a sea cucumber) Sticky Sea Cucumber Shellgrit Sea Cucumber Southern Sand Star (a sea cucumber) Dark-tipped Button Star Eight-armed Seastar Gunn‘s Six-armed Star Southern Six-armed Star Western Six-armed Star Eastern Six-armed Star Devil‘s Heart Urchin Large-plated Seastar Multi-spined Seastar Spotted Seastar (a sea star) The Saori Seastar Wilson‘s Seastar (a sea cucumber) (a sea cucumber) Troughton‘s Seastar (a sea star) (a brittle star) (a brittle star) Chequered Brittle Star (a brittle star) Ramsay‘s Serpent Star (a brittle star) (a brittle star) (a brittle star) Schayer‘s Brittle Star (a brittle star) Live-bearing Brittle Star Mottled Brittle Star (a brittle star) Spiny-armed Brittle Star (a brittle star) Large Spiny-armed Brittle Star Ophiura kinbergi (Ljungman, 1866) (orange basket star, in Ward et al., 2003) Paranepanthia grandis (Clark, 1928) Paracaudina australis (Semper, 1868) Paracaudina tetrapora (H.L. Clark, 1914) Parvulastra parvivipara (Keough & Dartnall, 1978) = Patiriella parvivipara Keough & Dartnall, 1978 Pentagonaster dubeni Gray, 1847 Pentocnus bursatus O'Loughlin & O'Hara, 1992 (recorded E and W of GAB; likely to occur in AW NRM) Peronella peronii (L. Agassiz, 1841) Petricia vernicina (Lamarck, 1816) Phyllacanthus irregularis Mortensen, 1928 Plectaster decanus (Müller & Troschel, 1843) Plesiocolochirus ignava (Ludwig, 1875) Protenaster australis (Gray, 1851) Pseudophidiaster rhysus H.L. Clark, 1916 Ptilometra australis (Wilton, 1843) (NB: A species from QLD and NSW; listed in Ward et al. 2003 as being recorded in the shelf waters of GAB) Ptilometra macronema (Müller, 1846) Ptilometra sp. (in Ward et al., 2003) Squamocnus aureoruber O'Loughlin & O'Hara, 1992 Staurothyone inconspicua (Bell, 1887) Stichopus ludwigi Erwe, 1913 Stichopus mollis (Hutton, 1872) = Australostichopus mollis (Hutton, 1872) (strawberry-surface holothurian, in Ward et al., 2003) Taeniogyrus roebucki (Joshua, 1914) Tosia australis Gray, 1840 / Tosia nobilis (Müller & Troschel, 1843) Temnopleurus michaelseni (Döderlein, 1914) Thyone okeni Bell, 1884 Thyone nigra Joshua & Creed, 1915 Trachythyone glebosa O'Loughlin & O'Hara, 1992 Uniophora granifera (Lamarck, 1816) (unidentified urchin, in Ward et al., 2003) (wide mouthed asteroid, in Ward et al., 2003) (yellow asteroid, in Ward et al., 2003) 143 Kinberg‘s Brittle Star (a basket star) (a sea star) Podgy Sea Cucumber (a sea cucumber) Little Patti; Pygmy Live-bearing Seastar Vermilion Biscuit Star (a sea cucumber) (a sand dollar) Cushion Star Western Slate-pencil Urchin Mosaic Seastar Orange-flecked Sea Cucumber (a heart urchin) (a sea star) (a feather star / crinoid) (a feather star / crinoid) (a faether star / crinoid) (a sea cucumber) (a sea cucumber) Ludwig‘s Sea Cucumber Southern Sea Cucumber (a sea cucumber) (a sea cucumber) Southern Biscuit Star (a sea urchin) (a sea cucumber) (a sea cucumber) (a sea cucumber) Granular Seastar (a sea urchin) (a sea star) (a sea star) Appendix 3: Shelled gastropod mollusc species known from the Great Australian Bight upper and mid shelf waters, and thus highly likely to occur in the AW NRM marine region (Cotton and Godfrey, 1931,1932,1938; Ponder and Yoo, 1976, 1977a,b, 1980; Bratcher and Cernohorsky, 1987; Ponder and de Keyzer, 1992; Henning and Hemmen, 1993; Wilson et al. 1993 and 1994; Bail and Limpus, 1997; Ponder and Grayson, 1998; Darragh, 2002; Ponder et al., 2002; Ponder, 2002, in ABRS, 2009; Watters, 2004; Academy of Natural Sciences, 2006; Australian Museum records, South Australian Museum records, Museum of Victoria records, Queen Victoria Museum and Art Gallery records, cited in OZCAM 2009). Common names as listed in Edgar (2008) and/or ABRS (2009). E = apparently endemic within South Australia. A = recorded directly east and/or west of AW NRM region, and therefore likely to occur in the region. I = introduced. * = taxonomy uncertain. Scientific Name Aclophora hedleyi Marshall, 1983 A Aclophoropsis festiva (A. Adams, 1851) A Aesopus australis (Angas, 1877) * Afrolittorina praetermissa (May, 1909) = Nodilittorina praetermissa (May, 1909) Alaba monile A. Adams, 1862 Alaba pulchra A. Adams, 1862 Alaba translucida (Hedley, 1905) Alaginella borda (Cotton, 1944) A Alaginella geminata (Hedley, 1912) Alaginella malina (Hedley, 1915) A Alocospira edithae (Pritchard & Gatliff, 1899) A Alocospira marginata (Lamarck, 1811) Alocospira oblonga (Sowerby, 1830) A Alocospira petterdi (Tate, 1893) A Alcyna acia Cotton, 1948 A Alvania (Alvania) novarensis Frauenfeld, 1867 A Alvania (Alvania) occidua (Cotton, 1944) A Alvania (Alvania) strangei (Brazier, 1894) A Alvania (Linemera) verconiana (Hedley, 1911) Amalda coccinata Kilburn, 1980 Amblychilepas javanicensis (Lamarck, 1822) A Amblychilepas nigrita (Sowerby, 1834) Amblychilepas oblonga (Menke, 1843) Amoria exoptanda (Reeve, 1849) Amoria undulata (Lamarck, 1804) Anabathron (Anabathron) contabulatum Frauenfeld, 1867 Anachis atkinsoni (Tenison Woods, 1876) Anachis beachportensis (Verco, 1910) Anachis cominellaeformis (Tate, 1892) Anachis dolicha (Verco, 1910) E, A Anachis fenestrata (Verco, 1910) E, A Anachis fulgida (Reeve, 1859) Anachis remoensis Gatliff & Gabriel, 1910 Anapella amygdala (Crosse & Fischer, 1864) Anapella cycladea (Lamarck, 1818) A Antisabia erma (Cotton, 1938) A Anatrophon latior (Verco, 1909) A Antiguraleus kingensis (Petterd, 1879) Antisabia erma (Cotton, 1938) A Common Name (a creeper shell) (a creeper shell) (a dove shell) Checkered Australwink (a small shell in the Litiopidae family) (a small shell in the Litiopidae family) (a small transparent shell in the Litiopidae family) (a small marginella shell from shallow shelf to upper slope) (a small marginella shell from shelf to upper slope) (a small marginella shell) Edith‘s Ancilla (a small ancilla shell) (a small ancilla shell) (a small ancilla shell) Tread Top Shell (a small rissoid gastropod that feeds on micro-algal film) (a small rissoid gastropod that feeds on micro-algal film) (a small rissoid gastropod that feeds on micro-algal film) (a small rissoid gastropod that feeds on micro-algal film) (an ancillid shell found on the continental shelf and slope) Rayed Keyhole Limpet / Saddle Keyhole Limpet / Javan Keyhole Limpet Black Keyhole Limpet (a keyhole limpet) Desirable Volute / Much-Desired Volute Wavy Volute (a gastropod in the Anabathridae family) (a small dove shell) (a small dove shell) (a small dove shell from the intertidal) (a small dove shell) (a small dove shell) (a small dove shell) (a small dove shell from the intertidal) (a small wedge shell from intertidal sand habitats) (a wedge shell from intertidal sand or mud) (a horse hoof limpet) (a small trophine shell) (a horse hoof limpet) 144 Antisabia foliacea (Quoy & Gaimard, 1835) Argalista corallina (Cotton & Godfrey, 1935) A Argobuccinum pustulosum tumidum (Dunker, 1862) A Ascorhis victoriae (Tenison-Woods, 1878) Assiminea (Metassiminea) brazier A Astele (Astele) armillatum (Wood, 1828) A Astele (Callistele) calliston Verco, 1905 A Astele (Astele) ciliare (Menke, 1843) Astele (Astele) rubiginosum (Valenciennes, 1846) A Astele (Astele) subcarinatum (Swainson, 1855) A Astele (Astelena) multigranum (Dunker, 1871) A Astele (Sinutor) incertum (Reeve, 1863) Asteracmea alboradiata (Verco, 1906) E Asteracmea crebristriata (Verco, 1904) Asteracmea illibrata (Verco, 1906) Asteracmea roseoradiata (Verco, 1912) A Asteracmea stowae (Verco, 1906) Asperdaphne (Asperdaphne) bastowi (Gatliff & Gabriel, 1908) A Asperdaphne (Asperdaphne) bitorquata (Sowerby, 1897) A Asperdaphne (Asperdaphne) desalesii (TenisonWoods, 1877) Asperdaphne (Asperdaphne) perplexa (Verco, 1909) A, E Asperdaphne (Asperdaphne) tasmanica (TenisonWoods, 1877) A Asperdaphne (Asperdaphne) vestalis (Hedley, 1903) A Asperdaphne (Asperdaphne) vercoi (Sowerby, 1897) A, E Asperdaphne (Asperdaphne) walcotae (Sowerby, 1893) A, E Asperdaphne (Aspertilla) legrandi (Beddome, 1883) A Astralium aureum (Jonas, 1844) Astralium rutidoloma (Tate, 1893) A, E Astralium squamiferum (Koch, 1844) Ataxocerithium beasleyi Cotton & Godfrey, 1938 A Austrocochlea constricta (Lamarck, 1822) Austrocochlea odontis (Wood, 1828) Austrocochlea porcata (A. Adams, 1853) Austrodrillia (Austrodrillia) agrestis (Verco, 1909) A, E Austrodrillia (Austrodrillia) dimidiata (Sowerby, 1897) A, E Austrodrillia (Austrodrillia) subplicata (Verco, 1909) A, E Austroharpa (Palamharpa) exquisita (Iredale, 1931) A Austroharpa (Palamharpa) punctata (Verco, 1896) Austroharpa (Unplaced) learorum Hart & Limpus, 1998 Austroliotia australis (Kiener, 1839) Austroliotia botanica (Hedley, 1915) A Austroliotia densilineata (Tate, 1899) Austroliotia pulcherrima (Reeve, 1843) (a horse hoof limpet common in intertidal & shallow subtidal, attached to underside of stones) Coral-Red Gibbula Flag Triton / Argus Triton (a hydrobiid snail) (a small brown snail) Jewelled Top Shell / Meyer's Top Shell Beautiful Top shell Calliope Top Shell Rusty Top Shell Subcarinate Astele / Keeled Top Shell / Umbilicated Top Shell (a top shell) Left-handed or Doubtful Calliostoma / Uncertain Top Shell White Rayed Sugar Limpet Fine Ridged Limpet Midget Limpet / Plain Limpet Rose Ray Limpet Stow's Limpet (a turrid shell) (a turrid shell) (a turrid shell) (a turrid shell) (a turrid shell) (a turrid shell) (a turrid shell) (a turrid shell) (a turrid shell) Golden Small Star Granular Small Star Star Shell / Scaly Star Shell (a small cerithiopsid shell that lives on sponges) Ribbed Top Shell Chequered Top Shell Zebra Top Shell (a turrid shell) (a turrid shell) (a turrid shell) Exquisite Harp Spotted Harp (a harp shell) (a liotine / turbinid shell) (a liotine / turbinid shell) (a liotine / turbinid shell) (a liotine / turbinid shell) 145 Austrolittorina unifasciata (Gray, 1826) Austromitra analogica (Reeve, 1845) Austromitra arnoldi (Verco, 1909) Austromitra minutenodosa Cernohorsky, 1980 Badepigrus pupoides (H. Adams, 1865) * Belloliva triticea (Duclos, 1835) A Bembicium auratum (Quoy & Gaimard, 1834) Bembicium vittatum Philippi, 1846 A Botelloides bassianus bassianus (Hedley, 1911) Botelloides bassianus borda Cotton, 1944 Botelloides chrysalidus chrysalidus (Chapman & Gabriel, 1914) Botelloides sulcatus sulcatus (Cotton, 1944) Bouchetriphora pallida (Pease, 1870) Buccinulum bednalli (Sowerby, 1895) Bulla quoyii Gray in Dieffenbach, 1843 Cabestana tabulata (Menke, 1843) A Cabestana spengleri (Perry, 1811) A Cacozeliana granarium (Kiener, 1842) A Cacozeliana icarus (Bayle, 1880) A Caecum (Caecum) amputatum Hedley, 1894 A Calliostoma (Fautor) allporti (Tenison Woods, 1876) A Calliostoma (Fautor) columnarium Hedley & May, 1908 A Calliostoma (Fautor) comptum (Adams, 1854) A Calliostoma (Fautor) hedleyi Pritchard & Gatliff, 1902 Calliostoma (Fautor) legrandi legrandi (TenisonWoods, 1876) Calliostoma (Fautor) zietzi Verco, 1905 A Calyptraea calyptraeformis (Lamarck, 1822) Cancellaria (Nevia) spirata Lamarck, 1822 Cancellaria (Sydaphera) undulata Sowerby, 1849 A Cantharidella beachportensis Cotton & Godfrey, 1934 A Cantharidella ocelllina (Hedley, 1911) A, E Cantharidus ramburi (Crosse, 1864) Capulus devotus Hedley, 1904 A Capulus violaceus Angas, 1867 Cardiolucina crassilirata A Cassis (Hypocassis) fimbriata Quoy & Gaimard, 1833 A Cellana solida (Blainville, 1825) A Cellana tramoserica (Holten, 1802) Charisma carinata (Verco, 1907) A, E Charisma josephi (Tenison Woods, 1877) A Cheilea flindersi Cotton, 1935 A Chevallieria australis Ponder, 1984 E Chicoreus (Triplex) damicornis (Hedley, 1903) Cingulina spina (Crosse & Fischer, 1864) Circulus delectabile (Tate, 1899) A Circulus harriettae (Petterd, 1884) Cirsonella weldii (Tenison-Woods, 1877) Clanculus albanyensis Jansen, 1995 A Clanculus consobrinus Tate, 1893 Clanculus denticulatus (Gray, 1827) A Blue Australwink (a small, variable costellate mitre shell found amongst rocks and macroalgae) (a small costellate mitre shell) (a gastropod in the Anabathridae family) (a small, common olivella shell) Gold-mouthed Conniwink (a littorinid shell found in a variety of nearshore habitats) (a top shell) (a top shell) (a top shell) (a top shell) (a small triphorid shell that feeds on sponges) (a buccinid whelk) Brown Bubble Shell / Brown Bubble Snail Ploughed Triton / Shouldered Triton / Tabulate Triton Spengler‘s Triton / Spengler‘s Rock Whelk (a small creeper shell) (a small creeper shell) (a small, tube-shaped gastropod in the Caecidae family) (a top shell found on the continental shelf and slope) Cape Pillar Top Shell (a top shell found on cup-shaped sponges) Hedley‘s Top Shell Legrand‘s Top Shell Zietz‘s Top Shell Shelf Limpet Spiral Nutmeg Shell (a nutmeg shell) Beachport Top Shell Eyelet Top Shell Rambur's Jewel Top Shell Devout Cap Shell Violet Cap Shell Densely Striated Lucina Fringed Helmet Shell / Fimbriate Helmet Shell Orange-edged Limpet Common Limpet Carinate Charisma / (a top shell) Joseph‘s Charisma (a limpet-like shell that attaches to stones or other shells) (a small gastropod in the Iravadiidae family) Damicornis Murex / Long-Horned Murex / Purple Murex (a pyramid shell) (a very small gastropod in the Vitrinellidae family) (a very small gastropod in the Vitrinellidae family) (a small shell in the Skeneidae family Yellow Top shell (a small top shell, common under stones on shallow reefs) (a small top shell from the intertidal and shallow subtidal) 146 Clanculus dunkeri (Koch in Philippi, 1843) A Clanculus euchelioides Tate, 1893 Clanculus flagellatus (Philippi, 1848) A Clanculus limbatus (Quoy & Gaimard, 1834) Clanculus maxillatus (Menke, 1843) Clanculus personatus (Philippi, 1849) A Clanculus philippi (Koch in Philippi, 1843) Clanculus plebejus (Philippi, 1852) Clanculus ringens (Menke, 1843) A Clanculus weedingi Cotton, 1938 A Codakia (Codakia) perobliqua (Tate, 1892) A Colpospira (Acutospira) accisa (Watson, 1881) Colpospira (Colpospira) mediolevis (Verco, 1910) Colpospira (Colpospira) runcinata (Watson, 1881) Colpospira (Colpospira) translucida Garrard, 1972 Cominella (Cominella) eburnea (Reeve, 1846) Cominella (Cominella) lineolata (Lamarck, 1809) Cominella (Godfreyna) torri Verco, 1909 Cominella (Josepha) tasmanica (Tenison Woods, 1878) Comitas murrawolga (Garrard, 1961) A Coralliophila (Coralliophila) mira (Cotton & Godfrey, 1932) Coralliophila (Coralliophila) wilsoni Prichard & Gatliff, 1898 A Cosmetalepas concatenatus (Crosse & Fischer, 1864) Crassispira (Crassispira) harpularia (Desmoulins, 1842) A Crassitoniella erratica erratica (May, 1913) Crepidula immersa (Angas, 1865) Cymatium (Monoplex) parthenopeum (von Salis, 1793) A Cylichnatys campanula Burn, 1978 Cystiscus angasi (Crosse, 1870) Cystiscus connectans (May, 1911) Cystiscus cratericula (Tate & May, 1900) Cystiscus subauriculata (May, 1916) A Daphnella (Daphnella) botanica Hedley, 1918 Daphnella (Daphnella) diluta Sowerby, 1897 A, E Daphnella (Daphnella) stiphra Verco, 1909 A, E Dentimargo allporti (Tenison Woods, 1876) A Dentimargo jaffa (Cotton, 1944) Dentimargo kemblensis (Hedley, 1903) Dentimargo lodderae (May, 1911) Dentimargo mayii (Tate & May, 1900) A Dermomurex (Dermomurex) angustus (Verco, 1895) Dermomurex (Dermomurex) goldsteini (Tenison Woods, 1876) Dermomurex (Viator) howletti A Diala megapicalis Ponder & Keyzer, 1992 Diala suturalis (A. Adams, 1853) (a small top shell from the intertidal and shallow subtidal) (a small top shell, common under stones on shallow reefs) (a small top shell, common under stones on shallow reefs) Keeled Clanculus Rounded Clanculus (a small, uncommon top shell from the intertidal) (a small top shell found on macroalgae in the subtidal) Plebian Top Shell / Plebian Clanculus (a small top shell found under stones on shallow reefs) (a small top shell found on macroalgae in the subtidal) Codakia Shell / Codakia (a screw shell) (a screw shell) (a screw shell) (a screw shell) Ribbed Cominella Shell Spotted Cominella Shell / Chequerboard Snail Torr‘s Whelk / Torr's Buccinum Whelk Tasmanian Buccinum Whelk (a turrid shell from deeper continental shelf waters) Coral Shell (a small, white or pale pink gastropod) (a small, orange, keyhole limpet with a white sculptured shell) (a common turrid shell) (a small gastropod that reproduces by direct development) Southern Slipper Limpet / Elongate Slipper Limpet Neapolitan Triton / Hairy Triton / Hairy Whelk / Giant Hairy Triton (a brown bubble shell, found in intertidal sand or sandy-mud) (a small marginella shell) (a small marginella shell) (a small marginella shell) (a small marginella shell) (a turrid shell from shallow water on rocky shores) (a turrid shell) (a turrid shell) (a small marginella shell) (a small marginella shell) (a small marginella shell) (a small marginella shell) (a small marginella shell) (a murex shell) Goldstein's Trophon / (a murex shell) (a murex shell; the only living species in the Dermomurex subgenus Viator known from southern Australia) (a small shell found amongst seagrass rhizomes & algal turf (a small shell found amongst rocks, seagrass rhizomes & algal turf 147 Diastoma melanioides (Reeve, 1849) Dicathais orbita (Gmelin, 1791) = Thais orbita Diodora lincolnensis Cotton, 1930 Dissona maccoyi (Tenison - Woods, 1878) Dolicholatirus spiceri (Tenison Woods, 1876) Domiporta strangei (Angas, 1867) Eatoniella (Eatoniella) depressa Ponder & Yoo, 1978 A Eatoniella (Eatoniella) exigua Ponder & Yoo, 1978 A Eatoniella (Eatoniella) fulva Ponder & Yoo, 1978 A Eatoniella (Eatoniella) juliae Ponder & Yoo, 1978 A Eatoniella (Eatoniella) melanochroma (Tate, 1899) Eatoniella (Eatoniella) puniceolinea Ponder & Yoo, 1978 Eatoniella (Eatoniella) taylorae Ponder & Yoo, 1978 Ericusa fulgetrum (Sowerby, 1825) A Ericusa papillosa (Swainson, 1822) Eatoniella (Eatoniella) atropurpurea (Frauenfeld, 1867) Eatonina (Eatonina) sanguinolenta Ponder & Yoo, 1980 Eatonina (Eatonina) shirleyae Ponder & Yoo, 1980 Eatoniopsis (Pilitonia) westralis Ponder & Yoo, 1980 A Ethminolia elveri Cotton & Godfrey, 1938 A Ethminolia vitiliginea (Menke, 1843) A Emarginula (Emarginula) candida A. Adams, 1852 A Emarginula (Emarginula) dilecta A. Adams, 1852 A Emarginula (Emarginula) patula Cotton, 1930 A Emarginula (Emarginula) subtilitexta Verco, 1908 A Emarginula (Subzeidora) devota Thiele, 1915 A Epidirona flindersi (Cotton & Godfrey, 1938) A Epidirona philipineri (Tenison Woods, 1877) Epitonium (Hyaloscala) friabile (Sowerby, 1844) A Epitonium (Hyaloscala) jukesianum (Forbes, 1852) A Epitonium (Laeviscala) tacitum (Iredale, 1936) Epitonium (Lamelliscala) godfreyi Cotton, 1938 A Epitonium (Lamelliscala) minorum (Iredale, 1936) A Eulima acutissima (Sowerby, 1866) A Eulima augur augur Angas, 1865 A Eulima augur broadbente (Cotton & Godfrey, 1932) A Eulima bivittata (Adams, H.& A., 1853) A Eunaticina albosutura Verco, 1909 A Eunaticina umbilicata (Quoy & Gaimard, 1833) A Euseila pileata Cotton, 1951 A Eutriphora armillata (Verco, 1909) A Eutriphora cana (Verco, 1909) Eutriphora pseudocana Marshall, 1983 Eutriphora tricolor (Laseron, 1954) A Exomilus pentagonalis (Verco, 1896) Favartia (Murexiella) brazieri (Angas, 1877) A Fossarina (Minopa) legrandi Petterd, 1879 A Fossarina (Fossarina) petterdi Crosse, 1870 A Fusus bednalli (Brazier, 1875) (a diastoma shell found in sand amongst seagrass, in the shallow subtidal) Cart-rut Shell / Cartrut Shell Port Lincoln Keyhole Limpet (a gastropod in the Ovulidae family) (a small spindle shell) (a mitre shell) (a small gastropod in the Eatoniellidae family) (a small gastropod in the Eatoniellidae family) (a small gastropod in the Eatoniellidae family) (a small gastropod in the Eatoniellidae family) (a small gastropod in the Eatoniellidae family) (a small gastropod in the Eatoniellidae family) (a small gastropod in the Eatoniellidae family) Lightning Volute Marbeled Volute / Papillose Volute (a small gastropod found amongst algae, under stones, and in crevices in intertidal and shallow subtidal) (a small gastropods found amongst algal turf and stones in intertidal and shallow subtidal) (a small gastropod found amongst algal turf in the intertidal) a small gastropod found amongst algal turf and debris in the lower intertidal and shallow subtidal) (a small top shell from the shallow subtidal) (a small top shell from the shallow subtidal) (a slit limpet commonly found in beach drift) (a slit limpet of broad geographic range and depth range) Flat-notched Limpet Delicate Notched Limpet (a slit limpet commonly found in beach drift) (a turrid shell) (a turrid shell) (a small wentletrap shell, from intertidal and shallow subtidal) (a small, common wentletrap shell) (a small wentletrap shell found in the intertidal) (a small wentletrap shell from the intertidal and shallow subtidal) Philippine Ladder Shell (a small eulimid shell, parasitic on echinoderms) (a small eulimid shell, parasitic on echinoderms) (a small eulimid shell, parasitic on echinoderms) a small eulimid shell, parasitic on echinoderms White Band Sand Snail (a small moon snail / sand snail) (a small cerithiopsid gastropod that feeds on sponges) (a small triphorid gastropod that feed on sponges) (a small triphorid gastropod that feed on sponges) (a small triphorid gastropod that feed on sponges) (a small triphorid gastropod that feed on sponges) (a small murex shell) Legrand‘s Top shell (a small top shell) Bednall's Colubraria / (a small dog whelk) 148 Fusus reticulatus (Adams, 1855) A Gracilispira lineata (Kiener, 1844) A Granata imbricata (Lamarck, 1816) A Granulina elliottae (Cotton, 1944) A Guraleus (Guraleus) pictus (Adams & Angas, 1864) Guraleus (Mitraguraleus) australis (Adams & Angas, 1864) A Haliotis cyclobates Péron, 1816 Haliotis laevigata Donovan, 1808 Haliotis roei Gray, 1826 Haliotis rubra Leach, 1814 (including form conicopora) Haliotis scalaris (Leach, 1814) Hedleytriphora basimacula Marshall, 1983 A Hedleytriphora elata (Thiele, 1930) A Hedleytriphora fasciata (Tenison Woods, 1879) Hedleytriphora scitula (A. Adams, 1851) A Hemitoma (Montfortia) subemarginata (Blainville, 1819) Herpetopoma annectans (Tate, 1893) A Herpetopoma aspersus (Philippi, 1846) A Herpetopoma fenestrata (Tate, 1893) Herpetopoma pumilio (Tate, 1893) Herpetopoma scabriuscula (Angas, 1867) A Herpetopoma vixumbilicata (Tate, 1893) Hexaplex conatus (McMichael, 1964) Hinea brasiliana (Lamarck, 1822) A Hipponix australis (Lamarck, 1819) Hydroginella columnaria (Hedley & May, 1908) Hydroginella tridentata (Tate, 1878) A Hydroginella vincentiana (Cotton, 1944) Hypermastus mucronatus (Sowerby, 1866) A Inella carinata Marshall, 1983 A Inella intercalaris Marshall, 1983 A Inella obliqua (May, 1915) A Inglisella fischeri (Adams, 1860) A Incisura (Scissurona) rosea (Hedley, 1904) Incisura (Scissurona) vincentiana (Cotton, 1945) A Isotriphora amethystina Marshall, 1983 Isotriphora aureovincta (Verco, 1910) A Isotriphora nivea (Verco, 1909) A Isotriphora vercoi Marshall, 1983 A Lamellaria ophione Gray, 1849 Laetifautor spinulosum (Tate, 1893) A Laevilitorina (Laevilitorina) johnstoni (Cotton, 1945) A Latitriphora latilirata (Verco, 1909) A Latirus pulleinei Verco, 1895 A Leiopyrga octona Tate, 1891 A Lepsiella (Bedeva) paivae (Crosse, 1864) Lepsiella (Lepsiella) flindersi (Adams & Angas, 1863) A Lepsiella (Lepsiella) vinosa (Lamarck, 1822) Lippistes helicoides (Gmelin, 1791) A Reticulated Little Buccinum / (a small dog whelk) (a small ancillid shell) Tiled False Ear Shell / Rounded False Ear Shell / Imbricated False Ear Shell (a small marginella shell from the continental shelf and slope) (a small turrid shell) (a small turrid shell) (an abalone) Greenlip Abalone Roe‘s Abalone Blacklip Abalone (an abalone) (a small triphorid gastropod that feeds on sponges) (a small triphorid gastropod that feeds on sponges) (a small triphorid gastropod that feeds on sponges) (a small triphorid gastropod that feeds on sponges) (a slit limpet) (a top shell from rocky habitats) (a top shell found under stones in intertidal & shallow subtidal) (a top shell) Dwarf Bead Shell Scurfy Bead Shell Spotted Bead Shell (a murex shell) Yellow-Coated Clusterwink / Jockiwink (a horse hoof limpet) (a small marginella shell) (a small marginella shell) (a small marginella shell) (a small eulimid shell, parasitic on echinoderms) (a small triphorid shell that feeds on sponges) (a small triphorid shell that feeds on sponges) (a small triphorid shell that feeds on sponges) (a nutmeg shell of uncertain distribution, in SA and southern WA) (a small slit shell) Gulf St. Vincent Slit Shell (a small triphorid shell that feeds on sponges) (a small triphorid shell that feeds on sponges) (a small triphorid shell that feeds on sponges) (a small triphorid shell that feeds on sponges) (a gastropod with a thin, translucent shell) Prickly Calliostoma / (a top shell) (a littorinid shell) (a small triphorid gastropod that feed on sponges) ( a spindle shell) (a top shell) (an oyster drill shell) Flinders' Lepsiella Wine-mouthed Lepsiella (a small tropical turbinate shell related to cap limpets) 149 Litozamia petterdi (Crosse, 1870) Livonia nodiplicata (Cox, 1910) Livonia roadnightae (McCoy, 1881) (a small trophine shell) Cotton‘s Volute Roadnight's Volute Lottia onychitis (Menke, 1843) Lottia mixta (Reeve, 1855) Lottia septiformis (Quoy & Gaimard, 1834) A Lucidestea muratensis (Cotton, 1944) A Lyria mitraeformis (Lamarck, 1811) Macroschisma producta Adams, 1850 A (a small acmaeid limpet found on rock platforms) (a small acmaeid limpet found on rock platforms) (a limpet found on limestone in the intertidal zone) (a small rissoid gastropod that feeds on micro-algal film) Mitre-shaped Lyria (a keyhole limpet found buried in sand, or under stones, or in crevices) (a small turrid shell) (a horse hoof llimpet) (a small turrid shell) (a small eulimid shell, parasitic on echinoderms) (a small eulimid shell, parasitic on echinoderms) (a small eulimid shell, parasitic on echinoderms) (a small eulimid shell, parasitic on echinoderms) (a small eulimid shell, parasitic on echinoderms) (a small eulimid shell, parasitic on echinoderms) Southern Baler Shell / Milton‘s Baler Shell (a small marginella shell) (a small marginella shell) (a small marginella shell) (a small spindle shell) Girdled Top Shell Coral Red Top Shell (a common mitre shell from rocky intertidal and subtidal habitats) (a mitre shell found under stones in the intertidal and shallow subtidal) Tapering Dove Shell (a small dove shell, common amongst algae & intertidal rocks (a small dove shell) Macteola anomala (Angas, 1877) Malluvium devotus (Hedley, 1904) A Marita compta (Adams & Angas, 1864) Melanella gradata (Cotton & Godfrey, 1932) A Melanella mayi (Tate, 1900) A Melanella murrayae (Cotton & Godfrey, 1932) A Melanella orthopleura (Tate, 1898) A Melanella planicincta (Cotton & Godfrey, 1932) A Melanella tenisoni (Tryon, 1886) A Melo miltonis (Griffith & Pidgeon, 1834) A Mesoginella caducocincta (May, 1916) Mesoginella pygmaeoides (Singleton, 1937) Mesoginella turbinata (Sowerby, 1846) Microcolus dunkeri (Jonas, 1844) Minolops cincta (Cotton & Godfrey, 1938) E Minolops corallina Cotton & Godfrey, 1935 Mitra (Mitra) carbonaria Swainson, 1822 A Mitra (Mitra) glabra Swainson, 1821 A Mitrella (Dentimitrella) acuminata (Menke, 1843) A Mitrella (Dentimitrella) austrina (Gaskoin, 1851) A Mitrella (Dentimitrella) dictua (Tenison Woods, 1878) A Mitrella (Dentimitrella) lincolnensis (Reeve, 1859) Mitrella (Dentimitrella) pulla Gaskoin, 1852 A Mitrella (Dentimitrella) semiconvexa (Lamarck, 1822) Mitrella (Zemitrella) purpureocincta (Verco, 1910) A Monophorus nigrofusca (A. Adams, 1851) Monstrotyphis yatesi (Crosse & Fischer, 1865) Montfortula rugosa (Quoy & Gaimard, 1834) Muricopsis planilirata (Reeve, 1845) A Muricopsis umbilicatus (Tenison Woods, 1876) A Naccula compressa (Verco, 1906) Notoacmea alta (Oliver, 1926) Naccula punctata (Quoy & Gaimard, 1834) Nannamoria guntheri (Smith, 1886) Nannamoria johnclarki (Bail & Limpus, 1997) A Nanula flindersi Cotton & Godfrey, 1935 Nassarius (Alectrion) glans particeps (Hedley, 1915) Nassarius (Hima) pauperus (Gould, 1850) Nassarius (Niotha) nigellus (Reeve, 1854) (a small, common dove shell that occurs in seagrass and sand habitats) Long Dove Mitre Shell Semiconvexa Dove Shell (a small dove shell) (a small triphorid shell that feeds on sponges) Yates‘ Typhis Rugose Slit Limpet / False Limpet (a small, common murex shell from intertidal and shallow subtidal) Umbilicated Murex Compressed Limpet Dotted Limpet Gunther‘s Volute (a small volute shell found in sand and rubble) (a small top shell from the continental shelf and slope) (Particeps Dog Whelk) (a small dog whelk shell occurring in sandy mud habitats) (a small dog whelk shell occurring in sandy mud habitats) 150 Nassarius (Niotha) pauperatus (Lamarck, 1822) A Nassarius (Zeuxis) pyrrhus (Menke, 1843) Natica sertata Menke, 1843 A Natica sticta Verco, 1909 A Neotrigonia bednalli (Verco, 1907) Nepotilla fenestrata (Verco, 1909) A Nerita (Melanerita) atramentosa Reeve, 1855 Niveria (Cleotrivia) globosa (Sowerby, 1832) Notoacmea alta (Oliver, 1926) Notoacmea flammea (Quoy and Gaimard, 1834) Notoacmea mayi (May, 1923) Notocochlis subcostata (Tenison Woods, 1878) A Notocypraea comptoni (Gray, 1847) A Notocypraea declivis (Sowerby, 1870) A Notocypraea piperita (Gray, 1825) Notogibbula bicarinata (Adams, 1854) Notogibbula lehmanni (Menke, 1843) = Prothalotia lehmanni (Menke, 1843) Notogibbula preissiana (Philippi, 1849) Notopeplum translucidum (Verco, 1896) A Nototriphora regina (Hedley, 1903) Nototriphora vestita Marshall, 1983 A Notovoluta kreuslerae (Angas, 1865) A Notovoluta verconis (Tate, 1892) Obesula albovittata (Hedley, 1903) A Obesula profundior Marshall, 1983 A Odontotrochus chlorostomus (Menke, 1843) Oliva australis Duclos, 1835 Onoba (Ovirissoa) rubicunda (Tate & May, 1900) Opalia (Opalia) australis (Lamarck, 1822) A Opalia (Dentiscala) granosa (Quoy & Gaimard, 1834) A Ovaginella ovulum (Sowerby, 1846) Ovaginella tenisoni (Pritchard, 1900) Ovaginella whani (Pritchard & Gatliff, 1900) Parastrophia (Parastrophia) cygnicollis (Hedley, 1904) A Parviterebra brazieri (Angas, 1875) A Parviterebra trilineata (Adams & Angas, 1864) A Patella (Scutellastra) chapmani Tenison Woods, 1876 Patella (Scutellastra) laticostata Blainville, 1825 A Patella (Scutellastra) peronii Blainville, 1825 Patelloida alticostata (Angas, 1865) Patelloida insignis (Menke, 1843) Patelloida latistrigata (Angas, 1865) Patelloida mufria (Hedley, 1915) Patelloida profunda (Deshayes, 1863) Patelloida profunda calamus (Crosse & Fischer, 1864) Peculator bacatus Cernohorsky, 1980 A Peculator porphyria (Verco, 1896) Penion mandarinus (Duclos, 1831) Pepta stricta (Hedley, 1907) A Persicula albomaculata (May , 1911) A (a small dog whelk shell found on a variety of subtrate types) (a small, common dog whelk) (a small moon snail) Spotted Sand Shell / a small moon snail) Brooch Shell (a turrid shell known from the continental shelf and slope) Black Nerite / Periwinkle (a bean cowrie) Tall Limpet Flamed Limpet May‘s Beetle Limpet (a small moon snail) Compton‘s Cowrie Sloping Cowrie / Speckled Cowrie Peppered Cowrie / Two-Coloured Piperita Cowrie Cox‘s Top Shell Many Coloured Top Shell / Lehmann's Top Shell Twin Keeled Top Shell (a small, translucent volute shell) (a small triphorid gastropod that feed on sponges) (a small triphorid gastropod that feed on sponges) Kreusler's Volute Verco‘s Volute (a small triphorid shell that feeds on sponges) (a small triphorid shell that feeds on sponges) (a top shell of variable form, colour & pattern; common in seagrasses) Southern Olive / Australian Olive (a rice shell) Austral Wentletrap / Southern Wentletrap Granose Wentletrap (a small marginella shell with a broad depth range) (a small marginella shell) (a small marginella shell) (a gastropod in the Caecidae family) (a small dove shell) Three-line Auger / a dove shell Chapman's Limpet / Eight-rayed Limpet Giant Limpet Peron's Limpet / Scaly Limpet / Scorched Limpet Tall-ribbed Limpet (a small limpet found under stones) Lateral-striped Limpet (a small limpet, often on shells of other gastropods) (a limpet) (a limpet) (a small volutomitrid shell) (a very small volutomitrid shell found on sandy substrates) Mandarin Penion Shell / Southern Siphon Whelk (a nutmeg shell) (a small gastropod resembling a marginella shell) 151 Persicula pulchella (Kiener, 1830) Phasianella australis (Gmelin, 1791) Phasianella ventricosa Swainson, 1822 Phasianotrochus apicinus (Menke, 1843) Phasianotrochus bellulus (Philippi, 1845) Phasianotrochus eximius (Perry, 1811) Phasianotrochus irisodontes (Quoy & Gaimard, 1834) Phenacolepas alboradiata (Verco, 1906) Phenacolepas calva (Verco, 1906) A Phycothais reticulata (Quoy & Gaimard, 1832) A Pisinna approxima (Petterd, 1884) A Pisinna costata (Hedley, 1911) Pisinna frenchiensis (Gatliff & Gabriel, 1908) A Pisinna varicifera relata (Cotton, 1944) Pisinna voorwindei Ponder & Yoo, 1976 Plaxiphora (Plaxiphora) albida (Blainville, 1825) Plesiotrochus monachus (Crosse & Fischer, 1864) Pleuroploca australasia (Perry, 1811) A Polinices (Conuber) conicus (Lamarck, 1822) A Primovula verconis (Cotton & Godfrey, 1932) A Pterochelus triformis (Reeve, 1845) Proterato (Cypraeerato) bimaculata (Tate, 1878) A Proterato (Sulcerato) denticulata (Pritchard & Gatliff, 1901) A Prothalotia flindersi (Fischer, 1878) A Prothalotia pulcherrimus (Wood, 1828) Prototyphis angasi (Crosse, 1863) Pseudamycla dermestoidea (Lamarck, 1822) A Pseudopisinna gregaria gregaria (Laseron, 1950) A Pusillina (Haurakia) mediolaevis Cotton, 1944 A Pyrene bidentata (Menke, 1843) A Ranella australasia australasia (Perry, 1811) A Retusa pygmaea (A. Adams, 1850) A Rissoella (Jeffreysilla) confusa umbilicata Ponder & Yoo, 1977 A Rissoella (Rissoella) vitrea Ponder & Yoo, 1977 A Rissoina (Rissoina) nivea A. Adams, 1853 A Sabia australis (= Hipponix australis) Sabia conica (= Hipponix conicus) (Schumacher, 1817) A Sassia (Austrotriton) subdistorta Lamarck, 1822 Sassia (Cymatiella) columnaria (Hedley & May, 1908) Sassia (Cymatiella) eburnea (Reeve, 1844) Sassia (Cymatiella) sexcostata (Tate, 1888) Sassia (Cymatiella) verrucosa (Reeve, 1844) Scissurella cyprina Cotton & Godfrey, 1938 A Scutus (Scutus) antipodes (Montfort, 1810) Flat-topped Margin Shell Australian Pheasant Shell / Pheasant Shell / Painted Lady Swollen Pheasant Shell / Painted Lady Pointed Kelp Shell Necklace Kelp Shell / Elegant Kelp Shell / Necklace Weed Shell Green Jewel Top Shell / True Kelp Shell Kelp Shell / Rainbow Kelp Shell / Green Necklace Shell (a sugar limpet) (a sugar limpet) (a small whelk shell occurring in seagrass beds and under rocks, in the intertidal) (a small gastropod found on algae, under rocks & stones) (a small gastropod from continental shelf waters) (a small gastropod found on algae under stones) (a small gastropod from the continental shelf) (a small gastropod from the continental shelf) White Plaxiphora Chiton (a small gastropod found in seagrass in the shallow subtidal) (a large tulip shell) / Australian Horse Conch / Australian Tulip Conical Moon Snail / Conical Sand Snail (a small ovulid shell) (a broadly distributed shell found in sand, and on rocks, near seagrass) (a small eratos shell that feeds on ascidians) (a small eratos shell that feeds on ascidians) Flinder Top Shell Crimson Lip Weed Shell / Beautiful Cantharidus Angas‘ Murex (a small dove shell) (a small gastropod found amongst algal turf in intertidal and shallow subtidal) (a small rissoid gastropod that feeds on micro-algal film) (a common dove shell associated with seagrass beds and brown macroalgae) Australian (Brown) Triton / Australasian Trumpet (a bubble shell) (a small gastropod found in lower intertidal & shallow subtidal) (a small gastropod found in lower intertidal & shallow subtidal) (a small rissoid gastropod that feeds on micro-algal film) (a horse hoof limpet that lives in groups on the shells of other gastropods) Conical Horse-Hoof / Bonnet Limpet Distorted Rock Triton (a small triton shell found in deeper shelf waters) Ivory Triton / Lesueur's Triton (a small triton shell found in the subtidal) Little Southern Triton Venus Slit Shell Elephant Snail / Roman Shield Shell / Duck‘s-bill Limpet 152 Seilarex verconis Cotton, 1951 A Semicassis (Antephalium) adcocki (Sowerby, 1896) A Semicassis (Antephalium) semigranosum (Lamarck, 1822) A Semicassis (Antephalium) sinuosum (Verco, 1904) A Semicassis (Semicassis) pyrum (Lamarck, 1822) Serpulorbis (Cladopoda) novaehollandiae (Chenu, 1843) Serpulorbis (Cladopoda) sipho (Lamarck, 1818) A Serrata mustelina (Angas, 1871) Sigapatella calyptraeformis (Lamarck, 1822) Siliquaria (Siliquaria) australis Quoy & Gaimard, 1834 Siliquaria (Pyxipoma) weldii (Tenison Woods, 1876) Sinezona atkinsoni (Tenison Woods, 1877) A Sinezona beddomei (Petterd, 1884) A Sinezona pacifica (Oliver, 1915) A Sinezona pulchra (Petterd, 1884) A Sinum zonale (Quoy & Gaimard, 1832) Siphonaria (Siphonaria) diemenensis Quoy & Gaimard, 1833 Siphonaria jeanae Jenkins, 1984 Siphonaria zelandica Quoy & Gaimard, 1833 Spectamen marsus Cotton & Godfrey, 1938 Specula regina Cotton, 1951 A Specula turbonilloides (Tenison Woods, 1879) A Spisula (Notospisula) trigonella (Lamarck, 1819) A Splendrillia (Splendrillia) woodsi (Beddome, 1883) Stosicia hedleyi (Tate, 1899) Stomatella auricula (Lamarck, 1816) Stomatella impertusa (Burrow, 1815) Tanea sagittata (Menke, 1843) Teretriphora spica (Verco, 1909) Tasmatica schoutanica (May, 1912) Tetraphora granifera (Brazier, 1894) Thalotia conica (Gray, 1827) A Tonna variegata (Lamarck, 1822) A Tricolia fordiana (Pilsbry, 1888) A Tricolia rosea (Angas, 1867) A Tricolia tomlini (Gatliff & Gabriel, 1921) A Tricolia variabilis (Pease, 1861) A Tripterotyphis robustus (Verco, 1895) Trivia (Ellatrivia) merces (Iredale, 1924) A Tugali cicatricosa A. Adams, 1852 A Turbo (Dinassovica) jourdani Kiener, 1839 Turbo (Euninella) gruneri Philippi, 1846 Turbo (Ninella) torquatus Gmelin, 1791 Turbo (Subninella) undulatus Lightfoot, 1786 Umbilica (Umbilica) armeniaca (Verco, 1912) Vasum (Altivasum) flindersi (Verco, 1914) A Vaceuchelus ampullus (Tate, 1893) A Vaceuchelus profundior (May, 1915) A Velacumantus australis (Quoy & Gaimard, 1834) Vexillum (Costellaria) acromiale (Hedley, 1915) A Vexillum (Costellaria) apicitinctum (Verco, 1896) (a small triphorid shell that feeds on sponges) (a helmet shell found on the continental shelf) Half-grained Helmet (a helmet shell found on the continental shelf) (a helmet shell found on the continental shelf and slope) (a worm shell that cements its shell to hard substrates) (a worm shell that cements its shell to hard substrates) (a small marginella shell from shallow reefs / rocky habitats) (a slipper shell) (a slit worm shell) (a slit worm shell) Atkinson‘s Slit Shell Beddome‘s Slit Shell (a small Scissurellid slit shell) Beautiful Slit Shell (a moon snail found on sandflats) Van Diemen‘s Siphon Shell (a limpet) New Zealand Siphon Shell Streaked Top Shell (a creeper shell that feeds on sponges) (a creeper shell that feeds on sponges) (a trough shell) (a turrid shell) (a small rissoid gastropod that feeds on micro-algal film) (a stomatella / false ear shell) False Ear Shell / Elongate False Ear Shell (a small moon snail found in intertidal zone) (a small triphorid gastropod that feed on sponges) (a small moon snail) (a small, common triphorid shell that feeds on sponges) Conical Thalotia Variegated Tun Shell (a small, tropical turbinid shell, common on rocky shores) Rosy Pheasant Shell Tomlin's Pheasant Shell (a small turbinid shell that is extremely variable in colour) Robust Pterynotus / Robust Typhis Common Southern Bean Cowry / Australian Bean Cowrie Scar False Limpet / Scarred Notched Limpet Jourdan‘s Turban Shell / Turban Shell Gruner's Turbo / Turban Shell Heavy Turban Shell / Turban Shell Common Warrener / Wavy Turban Apricot-Coloured Cowrie Flinders Vase Shell (a cream-coloured top shell) (a top shell from the continental shelf and slope) Southern Mud Creeper / Australian Mud Whelk (a costellate mitre shell) (a small costellate mitre shell) 153 Vexillum (Costellaria) lincolnense (Angas, 1878) A Vexillum (Pusia) australe (Swainson, 1820) Xenophora (Austrophora) flindersi flindersi (Cotton and Godfrey, 1938) Zaclys semilaevis (Tenison-Woods, 1877) Zaclys styliferus Cotton, 1951 A Zeacumantus diemenesis (Quoy & Gaimard, 1834) A Zeacumantus subcarinatus (Sowerby, 1855) A, I Zeidora legrandi Tate, 1894 A Zenepos mimica (Sowerby, 1897) A Zenepos minuta (Tenison-Woods, 1877) A Zoila friendii thersites (Gaskoin, 1849) A = Cypraea (Zoila) friendii thersites Zoila marginata orientalis / Zoila orientalis Zoila rosselli Cotton, 1948 A Zoila venusta (Sowerby, 1847) (a small costellate mitre shell found under rocks in intertidal and shallow subtidal) (a costellate mitre shell) (a carrier shell that attaches small shells & pebbles to its own shell (a small cerithiopsid gastropod that feeds on sponges) (a small cerithiopsid gastropod that feeds on sponges) (a mud creeper shell) (a mud creeper shell) (a slit limpet) (a turrid shell) (a turrid shell) Hump-backed Cowrie / Black Cowrie. Considered to be a geographically isolated eastern sub-species of the Z. friendii complex (Wilson et al., 1993; Wilson and Clarkson, 2004). Broad-margined Cowrie Rossell's Cowry (a cowrie with various named forms, at least one of which occurs in the SA portion of the Great Australian Bight) 154 Appendix 4: Infaunal species recorded in sediments at Stations 1, 2 ,3 , 4, 5 and 6 (inner and mid-shelf locations, closest to seaward of boundary of AW NRM region), during a 2006 SARDI survey (data from Currie et al., 2007). Genera in quotation marks may be misidentified. Major Taxonomic Group Crabs Carid shrimp / comb shrimps Craylets Snapping shrimps / pistol prawns Amphipod crustaceans Isopod crustaceans Mysid crustaceans Tanaid crustaceans Cumacean crustaceans Leptostracan crustaceans Ostracod crustaceans Stalked barnacles Bivalve molluscs Species Recorded Actaea peronii Halicarcinus rostratus Pisidia dispar (listed as Porcellana dispar) Leptochela sydniensis Philocheras intermedius Galathea australiensis Alpheus villosus Synalpheus fossor Birubius sp. Birubius drummondae Caprella scaura Ceradocus rubromaculatus Ceradocus serratus Cheiriphotis australiae Cyproidea sp. Dulichiella australis Gammaropsis (Gammaropsis) persetosus “ Halicreion” sp. Leucothoe diemenensis (listed as L. spinicarpa) Leucothoe sp. Mallacoota sp. Metaphoxus yaranellus Paradexamine echuca Urohaustorius halei Waldeckia kroyeri Waldeckia sp. Xenocheira fasciata Cerceis sp. Chitonopsis sp. Eurydice binda Haliophasma sp. Heteroserolis longicaudata (listed as Serolis longicaudata) Natatolana longispina Stenetrium armatum (species in Corophiidae family) Paranchialina angusta (species in Anarthruridae family) Kalliapseudes obtusifrons Paratanais sp. (listed as P. ignotus) Paratanais sp. Leptocuma pulleini Gynodiastylis truncatifrons Paranebalia longipes (species in Candonidae family) (species in Philomedidae family) Smilium peronii Carditella (Carditella) valida Donax (Tentidonax) francisensis Glycymeris (Glycymeris) striatularis Hiatella australis Mimachlamys asperrima Modiolus sp. Montacuta meridionalis Musculus nanus Neotrigonia bednalli Ovacuna atkinsoni (listed as Cuna solida) Sunetta vaginalis 155 Gastropod molluscs Nudibranchs Chitons Brittlestars Sand dollars Sea urchins Polychaete worms Sabellid worms (fan worms) Scale worms Nemertean worms Sipunculan worms Serpulid worms Cepahalochordates Bony fishes Oliva sp. 2 Oliva sp. 3 Retusa pygmaea Doris chrysoderma (listed as Neodoris chrysoderma) (more than one species in Chromodorididae) Leptochiton collusor (listed as Parachiton collusor) Ophiura kinbergi Ophiothrix (Ophiothrix) caespitosa Fibularia acuta Goniocidaris tubaria (species in Amphinomidae family) (species in Dorvilleidae family) (species in Glyceridae family) (species in Lumbrineridae family) (species in Nereididae family) (species in Onuphidae family) (species in Polygordiidae family) (2 species in Orbiniidae family) (2 species in Phyllodocidae family) (species in Polygordiidae family) (species in Syllidae family) (species in Spionidae family) ―Ampharete‖ sp. Eunice sp. Eurythoe complanata Exogone sp. Flabelligera sp. 1 ―Leiocapitella‖ sp. Lysidice sp. ―Micronephtys‖ sp. ―Nematonereis‖ sp. Notomastus sp. Ophelia sp. Pista sp. Syllis gracilis (a species in Sabellidae) (2 species in Polynoidae family) (2 species in Cephalothricidae family) Phascolosoma (Phascolosoma) annulatum (species in Serpulidae) Epigonichthys australis Foetorepus phasis 156 Appendix 5: Bony fish species likely to occur in the marine component of the AW NRM region (to 3 nautical miles from shore). V = oceanic species; rare inshore, as vagrant. E = apparently endemic within South Australia. P = Protected under the Fisheries Management (General) Regulations 2007 of the Fisheries Management Act 2007 in South Australia. U = might occur in AW NRM region, at upper depth limit. Compiled from: Kailola et al. (1993); Gomon et al. (1994; 2008); Kuiter (1996a, 1996b); Hutchins and Swainston (2001); SARDI survey data, identified by South Australian Museum (2002); R. Foster (S.A. Museum, pers. comm., 2006); M. Gomon (Museum Victoria, pers. comm., 2007); Australian Museum records, South Australian Museum records, Western Australian Museum records, Museum Victoria records, cited in OZCAM database (2009); Baker (2004, 2007, 2009); ABRS (2009); CSIRO (2009); Australian Museum (2009a). Scientific Name Common Name Acanthaluteres brownii (Richardson, 1846) Acanthaluteres spilomelanurus (Quoy & Gaimard, 1824) Acanthaluteres vittiger (Castelnau, 1873) Acanthistius serratus (Cuvier, 1828) Acanthopagrus butcheri (Munro, 1949) Acentronura australe Waite & Hale, 1921 = Idiotropiscis australe Whitley, 1947 Achoerodus gouldii (Richardson, 1843) Aetapcus maculatus (Günther, 1861) Afurcagobius tamarensis (= Favonigobius tamarensis) (Johnston, 1883) Alabes hoesei Springer & Fraser, 1976 Aldrichetta forsteri (Valenciennes, 1836) Allenichthys glauerti (Whitley, 1944) Allomycterus pilatus Whitley, 1931 Allothunnus fallai Serventy 1948 Ammotretis brevipinnis Norman, 1926 Ammotretis elongatus McCulloch, 1914 Ammotretis rostratus Günther, 1862 Anoplocapros amygdaloides Fraser-Brunner, 1941 = Anoplocapros robustus (Fraser-Brunner, 1941) Anoplocapros lenticularis (Richardson, 1841) Spiny-tailed Leatherjacket / Spinytail Leatherjacket Bridled Leatherjacket Toothbrush Leatherjacket Western Wirrah Black Bream / Bream Southern Little Pipehorse / Southern Pygmy Pipehorse Western Blue Groper / Blue Groper Warty Prowfish Tamar River Goby / Tamar Goby Dwarf Shore-Eel Yellow-eye Mullet / Yelloweye Mullet Glauert‘s Anglerfish Australian Burrfish / Small-spined Porcupine Fish Slender Tuna Short-fin Flounder / Shortfin Flounder Elongate Flounder Long-snout Flounder / Longsnout Flounder / Longsnouted Flounder / Bay Flounder Western Smooth Boxfish / Blue Boxfish / Robust Boxfish / Chubby Basketfish Aploactisoma milesii (Richardson, 1850) Apogonops anomalus Ogilby, 1896 Aracana aurita (Shaw, 1798) Aracana ornata (Gray, 1838) Argentina australiae Cohen, 1958 Humpback Boxfish / White-barred Boxfish / Humpty Dumpty Velvetfish Threespine Cardinalfish Shaw‘s Cowfish / Striped Cowfish Ornate Cowfish Silverside Arenogobius bifrenatus (Kner 1865) Bridled Goby Argyrosomus japonicas (Temminck & Schlegel, 1844) = Argyrosomus hololepidotus (Lacepède, 1801) Mulloway / Butterfish / Jewfish Arnoglossus muelleri (Klunzinger, 1872) Arnoglossus micrommatus Amaoka, Arai & Gomon 1997 Arripis georgianus = Arripis georgianus (Valenciennes, 1831) Arripis truttaceus = Arripis truttacea (Cuvier, 1829) Mueller‘s Flounder Smalleye Flounder / Flimsy flounder Australian Herring / Tommy Ruff Aseraggodes haackeanus (Steindachner, 1883) Aspasmogaster liorhyncha Briggs, 1955 Aspasmogaster tasmaniensis (Günther, 1861) Austrolabrus maculatus (Macleay, 1881) 157 Australian Salmon / West Australian Salmon / Western Australian Salmon Southern Sole / Southern Textile Sole Smooth-Snout Clingfish Tasmanian Clingfish Black-spotted Wrasse / Blackspotted Wrasse Auxis thazard (Lacépède , 1800) Beliops xanthokrossos Hardy, 1985 Bodianus frenchii (Klunzinger, 1880) Bodianus vulpinus (Richardson, 1850) Brachaluteres jacksonianus (Quoy & Gaimard, 1824) Brachionichthys australis Last, Gledhill and Holmes 2007 Brachynectes fasciatus Scott, 1957 Caesioperca lepidoptera (Forster, 1801) Caesioperca rasor (Richardson, 1839) Callanthias australis Ogilby, 1899 Callogobius depressus (Ramsay & Ogilby, 1886) Callogobius mucosus (Günther, 1872) Campichthys galei (Duncker, 1909) P Cantheschenia longipinnis (Fraser-Brunner, 1941) Caprichthys gymnura McCulloch & Waite, 1915 Capropygia unistriata (Kaup, 1855) Centroberyx australis Shimizu & Hutchins, 1987 Centroberyx gerrardi (Günther, 1887) Centroberyx lineatus (Cuvier, 1829) Centropogon latifrons Mees, 1862 Cepola australis Ogilby, 1889 Cheilodactylus nigripes Richardson, 1850 Cheilodactylus rubrolabiatus Allen & Heemstra, 1976 Chelidonichthys kumu (Cuvier, 1829) Chelmonops curiosus Kuiter, 1986 Chironemus georgianus Cuvier, 1829 Chironemus maculosus (Richardson, 1850) Chrysophrys auratus (Bloch & Schneider, 1801) (= Pagrus auratus) Cleidopus gloriamaris De Vis, 1882 Cnidoglanis macrocephalus (Valenciennes, 1840) Cochleoceps bicolor Hutchins, 1991 Bullet Mackerel Southern Longfin Western Foxfish / Foxfish Western Blackspot Pigfish Southern Pygmy Leatherjacket / Pygmy Leatherjacket Australian Handfish / Common Handfish Weedy Threefin / Southern Barred Threefin / Southern Barred Triplefin Butterfly Perch Barber Perch Splendid Perch / Rosy Perch Flathead Goby Sculptured Goby Gales Pipefish Smoothspine Leatherjacket Rigid Boxfish Spiny Boxfish Yellow-eyed Red Snapper / Yellow-eyed Nannygai Red Snapper / Redfish / Bight Redfish Swallowtail Western Fortescue Bandfish Magpie Perch / Black-striped Morwong Red-lipped Morwong / Redlip Morwong Red Gurnard / Flying Gurnard Talma / Western Talma / Square-back Butterflyfish / Truncate Coralfish Western Kelpfish / Southern Kelpfish / asselled Kelpfish Silverspot / Silver Spot Snapper / Pink Snapper Pineapple Fish / Pineapplefish / Knight Fish Estuary Catfish / Cobbler / Southern Cobbler Western Cleaner-Clingfish / Western Cleaner Clingfish Spade-nosed Clingfish / Spade-nose Clingfish Short-finned Conger Eel / Short-finned Conger Prickly Toadfish Barred Toadfish / Prickly Toadfish Common Dolphinfish Pink Sandfish Slender Sand-diver Southern Crested Weedfish / Crested Weedfish Southern Tongue Sole / Broadhurst‘s Tongue Sole Silver Dory Dusky Morwong / Butterfish Australian Tusk / Tusk Slender Blindfish Long-finned Pike Globe Fish / Porcupine Fish / Slender-spined Porcupine Fish Castelnau‘s Wrasse / Pretty Polly Sharksucker / Slender Suckerfish Prickly Anglerfish Sponge Anglerfish Cochleoceps spatula (Günther, 1861) Conger wilsoni (Bloch & Schneider, 1801) Contusus brevicaudus Hardy, 1981 Contusus richei (Fréminville, 1813) Coryphaena hippurus Linnaeus, 1758 Crapatalus munroi Last & Edgar 1987 Creedia haswelli (Ramsay, 1881) Cristiceps australis Valenciennes, 1836 Cynoglossus broadhursti Waite, 1905 Cyttus australis (Richardson, 1843) Dactylophora nigricans (Richardson, 1850) Dannevigia tusca Whitley, 1941 U Dermatopsis multiradiatus McCulloch & Waite, 1918 Dinolestes lewini (Griffith & Smith, 1834) Diodon nicthemerus Cuvier, 1818 Dotalabrus aurantiacus (Castelnau, 1872) Echeneis naucrates Linnaeus, 1758 Echinophryne crassispina McCulloch & Waite, 1918 Echinophryne reynoldsi Pietsch & Kuiter, 1984 158 Eeyorius hutchinsi Paulin, 1986 Emmelichthys nitidus Richardson, 1845 Engraulis australis (White, 1790) Enigmapercis reducta Whitley, 1936 Enoplosus armatus (White, 1790) Eocallionymus papilio (Günther, 1864) Epigonichthys australis (Raff, 1912) Etrumeus teres (De Kay, 1842) Eubalichthys bucephalus (Whitley, 1931) Eubalichthys cyanoura Hutchins, 1987 Eubalichthys mosaicus (Ramsay & Ogilby, 1886) Eubalichthys quadrispinis Hutchins 1977 (in deeper shelf waters ) Eupetrichthys angustipes Ramsay & Ogilby, 1888 Eviota bimaculata Lachner & Karnella, 1980 Favonigobius lateralis (Macleay, 1881) Fistularia petimba Lacepède, 1803 Foetorepus calauropomus (Richardson, 1844) Genus A, sp. 2 (Hutchins, in Gomon et al., 1994) Genus B sp. (Hutchins, in Gomon et al., 1994) Genypterus blacodes (Forster, 1801) Genypterus tigerinus Klunzinger, 1872 Girella zebra (Richardson, 1846) Glyptauchen panduratus (Richardson, 1850) Gnathanacanthus goetzeei Bleeker, 1855 Gnathophis longicauda = G. longicaudus or G. longicaudatus (Ramsay & Ogilby, 1888) Gnathophis umbrellabia (Whitley, 1946) Gonorynchus greyi (Richardson, 1845) Gymnapistes marmoratus (Cuvier, 1829) Gymnothorax prasinus (Richardson, 1848) Haletta semifasciata (Valenciennes, 1840) Finetooth Beardie Redbait Australian Anchovy Broad Duckbill / Broad Sandfish / Broad Sand-diver Old Wife Painted Stinkfish / Painted Dragonet Deepwater Lancelet Maray Black Reef Leatherjacket Blue-tailed Leatherjacket / Bluetail Leatherjacket Mosaic Leatherjacket Fourspine Leatherjacket Snakeskin Wrasse Twospot Fringedfin Goby / Two-spot Fringed-fin Goby / Twospot Goby Longfin Goby / Long-finned Goby / Spotted Goby Rough Flutemouth / Flutemouth Common Stinkfish Brown-spotted Spiny Clingfish / Kelp Clingfish Rat Clingfish Pink Ling Rock Ling Zebra Fish / Zebrafish Goblin Fish / Goblinfish Red Velvetfish Little Conger-Eel / Silver Conger / Little Conger Umbrella Conger Beaked Salmon Cobbler / South Australian Cobbler / Soldierfish Green Moray / Yellow Moray / Brown Reef Eel Weedy Whiting / Blue Weed Whiting / Blue Rock Whiting Yellowback Triplefin / Yellowback Threefin / Blackthroated Triplefin / Black-throated Threefin Reef Ocean Perch / Ocean Perch / Red Ocean Perch Western Upside-down Pipefish / Western Upside Down Pipefish Helcogramma decurrens McCulloch & Waite, 1918 Helicolenus percoides (Richardson & Solander, 1842) Heraldia sp. 1 = southern form of Heraldia nocturna Paxton, 1975 P Heteroclinus adelaidae Castelnau, 1872 Heteroclinus eckloniae (McKay, 1970) Heteroclinus heptaeolus (Ogilby, 1885) Heteroclinus macrophthalmus Hoese, 1976 Heteroclinus roseus (Günther, 1861) Heteroclinus sp. 5 (in Gomon et al., 1994) Heteroclinus sp. 6 (in Gomon et al., 1994) Heteroscarus acroptilus (Richardson, 1846) Hippocampus bleekeri Fowler, 1908 (N.B. H. bleekeri is Adelaide‘s Weedfish Kelp Weedfish Seven-Bar Weedfish / Sevenbar Weedfish Tasselled Weedfish / Large-Eye Weedfish Rosy Weedfish Fewray Weedfish Milward‘s Weedfish Rainbow Cale / Rainbowfish Southern Potbelly Seahorse / Pot-bellied Seahorse closely related to H. abdominalis Lesson, 1827) Hippocampus breviceps Peters, 1869 (in GAB there may be a western form of H. breviceps; in WA this has been assigned species status by some, and known as H. tuberculatus) Histiogamphelus cristatus (Macleay, 1881) P Histiophryne bougainvilli (Valenciennes, 1837) Histiophryne cryptacanthus (= H. cryptacantha) (Weber, 1913) 159 Short-headed Seahorse / Short-snouted Seahorse Macleay‘s Crested Pipefish / Rhino Pipefish Bougainville‘s Anglerfish / Smooth Anglerfish Rodless Anglerfish / Cryptic Anglerfish Hyperlophus vittatus (Castelnau, 1875) Hyperoglyphe antarctica (Carmichael, 1818) Hyporhamphus melanochir (Valenciennes, 1847) Hypoplectrodes nigroruber (= Hypoplectrodes nigrorubrum) (Cuvier, 1828) Ichthyoscopus barbatus Mees, 1960 Kathetostoma canaster Gomon & Last, 1987 Kathetostoma laeve (Bloch & Schneider, 1801) Kathetostoma nigrofasciatum (Waite & McCulloch, 1915) Katsuwonus pelamis (Linnaeus, 1758) Kaupus costatus (Waite and Hale, 1921) Kuiterichthys furcipilis (Cuvier, 1817) = Kuiterichthys sp. (western form) Kyphosus sydneyanus (Günther, 1886) Lagocephalus lagocephalus (Linnaeus, 1758) Lagocephalus sceleratus (Gmelin, 1789) Lampris guttatus (Brünnich, 1788) V Latris lineata (Forster, in Bloch & Schneider, 1801) Latropiscis purpurissatus (Richardson, 1843) Lepidoblennius marmoratus (Macleay, 1878) Lepidoperca filamenta Roberts, 1987 Lepidoperca occidentalis Whitley, 1951 Lepidopus caudatus (Euphrasen, 1788) V Lepidotrigla papilio (Cuvier, 1829) Lepidotrigla spinosa Gomon, 1987 Lepidotrigla vanessa (Richardson, 1839) Leptoichthys fistularius Kaup, 1853 P Lesueurina platycephala Fowler, 1908 Leviprora inops (Jenyns, 1840) Lissocampus caudalis Waite & Hale, 1921 P Lissocampus runa (Whitley, 1931) P Liza argentea (Quoy & Gaimard, 1825) Lophonectes gallus Günther, 1880 Lotella rhacina (= Lotella rhacinus) (Forster, in Bloch & Schneider, 1801) Macroramphosus scolopax (Linnaeus, 1758) Makaira indica (Cuvier, 1832) O Maroubra perserrata Whitley, 1948 P Masturus lanceolatus (Lienard, 1840) V Maxillicosta scabriceps Whitley, 1935 Maxillicosta meridianus Motomura, Last & Gomon, 2006 Metavelifera multiradiatus (Regan, 1907) Meuschenia flavolineata Hutchins, 1977 Meuschenia freycineti (Quoy & Gaimard, 1824) Meuschenia galii (Waite, 1905) Meuschenia hippocrepis (Quoy & Gaimard, 1824) Meuschenia scaber (Forster, 1801) Meuschenia venusta Hutchins, 1977 Mola mola (Linnaeus 1758) V Mola ramsayi (Giglioli 1883) V 160 Sandy Sprat Blue-eye Trevalla Sea Garfish / Southern Sea Garfish Black-banded Seaperch / Banded Seaperch Fringed Stargazer Speckled Stargazer Common Stargazer / Eastern Stargazer Deepwater Stargazer Skipjack Tuna Deepbody / Deep-bodied Pipefish Rough Anglerfish Silver Drummer / Southern Silver Drummer Ocean Puffer / Oceanic Pufferfish Giant Toado / Silver Toadfish / Silver Pufferfish / North-west Blowfish Opah Striped Trumpeter / Stripey Trumpeter / Tasmanian Trumpeter Sergeant Baker Jumping Blenny Western Orange Perch Slender Orange Perch Frostfish Spiny Gurnard / Southern Spiny Gurnard Southern Shortfin Gurnard Butterfly Gurnard Brushtail Pipefish / Brush-tail Pipefish Flathead Sandfish / Flathead Pygmy Stargazer Long-head Flathead / Longhead Flathead Smooth Pipefish Javelin Pipefish Jumping Mullet / Jumper Mullet / Flat-tail Mullet / Flattail Mullet / Goldspot Mullet Crested Flounder Beardie / Large-tooth Beardie Common Bellowsfish Black Marlin Sawtooth Pipefish Sharptail Sunfish Little Scorpionfish Southern Gurnard Perch Common Veilfin Yellow-striped Leatherjacket / Yellowstriped Leatherjacket / Yellow-tail Leatherjacket Six-spine Leatherjacket / Sixspine Leatherjacket / Six-spined Leatherjacket Blue-lined Leatherjacket / Bluelined Leatherjacket Horseshoe Leatherjacket Velvet Leatherjacket / Cosmopolitan Leatherjacket Stars and Stripes Leatherjacket / Stars-and-Stripes Leatherjacket Ocean Sunfish Short Sunfish Myxus elongatus Günther, 1861 Naucrates ductor (Linnaeus, 1758) Neatypus obliquus Waite, 1905 Nelusetta ayraudi (Quoy & Gaimard, 1824) Sand Mullet Pilotfish Western Footballer / Footballer Sweep Ocean Leatherjacket / Ocean Jacket / Chinaman Leatherjacket Jackass Morwong / Jackass Fish Nemadactylus macropterus (Forster, in Bloch & Schneider, 1801) Nemadactylus valenciennesi (Whitley, 1937) Queen Snapper / Blue Morwong / Southern Blue Morwong Little Weed Whiting / Little Rock Whiting Whiskered Prowfish Gulf Gurnard Perch Black-spotted Gurnard Perch Gurnard Perch / Bighead Gurnard Perch Common Gurnard Perch / Ruddy Gurnard Perch Thetis Fish Sailfin Goby / Castelnau‘s Goby Girdled Goby Groove-cheek Goby / Grooved-cheek Goby Groovecheek Goby / Groovecheeked Goby Red Pipefish Orange-spotted Wrasse / Brown-spotted Wrasse Herring Cale Ringed Toadfish / Ringed Pufferfish Blue-spotted Pufferfish / Blue-spotted Toadfish Variegated Snake-Blenny / Variegated Snakeblenny Adelaide Blenny / Adelaide Snake-Blenny / Adelaide Snakeblenny Short-finned Snake-Blenny / Shortfin Snakeblenny Black-Backed Snake-Blenny / Blackback SnakeBlenny / Blackback Snakeblenny Variable Snake-Blenny / Variable Snakeblenny Serpent Eel / Giant Snake Eel / Snake Eel Maori Wrasse Neoodax balteatus (Valenciennes, 1840) Neopataecus waterhousii (Castelnau, 1872) Neosebastes bougainvillii (Cuvier, 1829) Neosebastes nigropunctatus McCulloch, 1915 Neosebastes pandus (Richardson, 1842) Neosebastes scorpaenoides Guichenot, 1867 Neosebastes thetidis (Waite, 1899) Nesogobius pulchellus (Castelnau, 1872) Nesogobius maccullochi Hoese and Larson 2006 Nesogobius sp. 4 (in Gomon et al., 2008) Notiocampus ruber (Ramsay & Ogilby, 1886) P Notolabrus parilus (Richardson, 1850) Olisthops cyanomelas Richardson, 1850) Omegophora armilla (Waite & McCulloch, 1915) Omegophora cyanopunctata Hardy & Hutchins, 1981 Ophiclinops varius (McCulloch & Waite, 1918) Ophiclinus antarcticus Castelnau, 1872 Ophiclinus brevipinnis George & Springer, 1980 Ophiclinus gracilis (Waite, 1906) Ophiclinus ningulus George & Springer, 1980 Ophisurus serpens (Linnaeus, 1758) Ophthalmolepis lineolata (Valenciennes, 1838) = Ophthalmolepis lineolatus Oplegnathus woodwardi Waite 1900 Optivus agrammus Gomon, 2004 Othos dentex (Cuvier, 1828) Parablennius tasmanianus (Richardson, 1842) Parapercis sp. Parapercis haackei (Steindachner, 1884) Parapercis naevosa Serventy, 1937 Parapercis ramsayi Steindachner, 1884 Paraplesiops meleagris (Peters, 1869) Parapriacanthus elongatus (McCulloch, 1911) Paraulopus nigripinnis (Günther, 1878) Paratrachichthys macleayi (Johnston, 1881) Parazanclistius hutchinsi Hardy, 1983 Parequula melbournensis (Castelnau, 1872) Knifejaw Western Roughy Harlequin Fish Tasmanian Blenny Western Barrded Grubfish Wavy Grubfish Western Red-banded Grubfish Spotted Grubfish Western Blue Devil / Blue Devilfish Slender Bullseye Blacktip Cucumberfish Sandpaper Fish Short Boarfish / Hutchins‘ Boarfish Silverbelly / Southern Silverbelly / Melbourne Silverbelly / Melbourne Silver Biddy Yellow-spotted Boarfish / Brown-spotted Boarfish Giant Boarfish Victorian Scalyfin / Scalyfin Smallfin Clingfish / Little Clingfish Long-Snout Clingfish Obscure Clingfish / Obscure Little Clingfish Paristiopterus gallipavo Whitley, 1944 Paristiopterus labiosus (Günther, 1871) Parma victoriae (Günther, 1863) Parvicrepis parvipinnis (Waite, 1906) Parvicrepis sp. 1 (in Gomon et al., 2008) Parvicrepis sp. 2 (in Gomon et al., 2008) 161 Pegasus lancifer Kaup, 1861 = Acanthopegasus lancifer McCulloch, 1915 Pelates octolineatus (Jenyns, 1840) Pelsartia humeralis (Ogilby, 1899) Pempheris klunzingeri McCulloch, 1911 Pempheris multiradiata = Pempheris multiradiatus Klunzinger, 1880 Pempheris ornata Mooi & Jubb, 1996 Pentaceropsis recurvirostris (Richardson, 1845) Peronedys anguillaris Steindachner, 1883 Perryena leucometopon (Waite, 1922) Phycodurus eques (Günther, 1865) P Phyllophryne scortea (McCulloch & Waite, 1918) Phyllopteryx taeniolatus (Lacépède, 1804) P Pictilabrus laticlavius (Richardson, 1840) Plagiogeneion macrolepis McCulloch, 1914 Platycephalus aurimaculatus Knapp, 1987 Platycephalus bassensis Cuvier, 1829 Platycephalus conatus Waite & McCulloch, 1915 Platycephalus laevigatus = Leviprora laevigatus Cuvier & Valenciennes, 1829 Platycephalus longispinis Macleay, 1884 Platycephalus speculator Klunzinger, 1872 Polyprion oxygeneios (Forster, 1801) Polyspina piosae (Whitley, 1955) Pomatomus saltatrix Linnaeus, 1766 Posidonichthys hutchinsi Briggs, 1993 Pseudocaranx georgianus (Cuvier, 1833) Pseudocaranx wrighti (Whitley, 1931) Pseudogobius olorum (Sauvage, 1880) Pseudolabrus rubicundus (Macleay, 1881) = Pseudolabrus psittaculus (Richardson, 1840) Pseudophycis barbata Günther, 1863 Pseudophycis breviuscula (Richardson, 1846) Pseudorhombus arsius (Hamilton, 1822) Pseudorhombus jenynsii (Bleeker, 1855) Pterygotrigla polyommata (Richardson, 1839) Pugnaso curtirostris (Castelnau, 1872) P Ranzania laevis (Pennant 1776) V Remora australis (Bennett, 1840) Remora remora (Linnaeus, 1758) Repomucenus calcaratus (Macleay, 1881) = Callionymus calcaratus Rexea solandri (Cuvier 1832) Rhombosolea tapirina Günther, 1862 Rhycherus filamentosus (Castelnau, 1872) Rhycherus gloveri Pietsch, 1984 Sarda orientalis (Temminck & Schlegel, 1844) Sardinops sagax (Jenyns, 1842) Satyrichthys sp. 162 Sculptured Seamoth / Sculptured Sea Moth Western Striped Grunter / Striped Perch / Striped Trumpeter / Shitty Sea Trumpeter Rough Bullseye Common Bullseye Orange-lined Bullseye Long-snouted Boarfish / Long-snout Boarfish Eelblenny / Eel-blenny White-nose Pigfish / Whitenose Pigfish Leafy Seadragon White-spotted Anglerfish / Smooth Anglerfish Weedy Seadragon Senator Wrasse Bigscale Rubyfish Toothy Flathead Sand Flathead / Southern Sand Flathead Deepwater Flathead Grassy Flathead / Rock Flathead Long-spined Flathead Yank Flathead / Southern Blue-spotted Flathead Hapuku Orange-barred Pufferfish / Orangebarred Pufferfish Tailor Spiny Clingfish / Posidonia Clingfish Silver Trevally / White Trevally / Skipjack Trevally Sand Trevally / Skipjack Trevally Blue-spot Goby / Bluespot Goby / Swan River Goby Rosy Wrasse Bearded Rock Cod / Bearded Cod Bastard Red Cod Large-tooth Flounder / Largetooth Flounder / Largetoothed Flounder Small-tooth Flounder / Smalltooth Flounder / Smalltoothed Flounder Sharp-beaked Gurnard / Latchet Pug-nose Pipefish / Pugnose Pipefish Slender Sunfish Whalesucker Remora / Short Suckerfish Spotted Dragonet / Spotted Stinkfish Gemfish (western stock) Greenback Flounder / Melbourne Flounder / Southern Flounder Tasselled Anglerfish Glover‘s Anglerfish Oriental Bonito / Striped Bonito Australian Pilchard Southern Armour Gurnard Saurida undosquamis (Richardson, 1848) Scalanago lateralis Whitley, 1935 Schuettea woodwardi (Waite, 1905) Scobinichthys granulatus (Shaw, 1790) Scolecenchelys australis = Muraenichthys australis (Macleay, 1881) Scolecenchelys breviceps = Muraenichthys breviceps (Günther, 1876) Scomber australasicus Cuvier 1832 Scomberesox saurus (Walbaum, 1792) Scorpaena papillose (Bloch & Schneider, 1801) Scorpis aequipinnis Richardson, 1848 Scorpis georgianus (= Scorpis georgiana) Valenciennes, 1832 Seriola hippos Günther, 1876 Seriola lalandi Valenciennes, 1833 Seriolella brama (Günther, 1860) Seriolella punctata (Forster, 1801) Sillaginodes punctata (= Sillaginodes punctatus) (Cuvier, 1829) Sillago bassensis Cuvier, 1829 Sillago schomburgkii Peters, 1864 Siphamia cephalotes (Castelnau, 1875) = Siphaemia cephalotes Siphonognathus argyrophanes Richardson, 1858 Siphonognathus attenuatus (Ogilby, 1897) Siphonognathus beddomei (Johnston, 1885) Siphonognathus caninus (Scott, 1976) Siphonognathus radiatus (Quoy & Gaimard, 1834) Siphonognathus tanyourus Gomon & Paxton, 1986 Sphyraena novaehollandiae Günther, 1860 Spratelloides robustus Ogilby, 1897 Sticharium clarkae George & Springer, 1980 Sticharium dorsale Günther, 1867 Stigmatopora argus (Richardson, 1840) P Stigmatopora nigra Kaup, 1856 P Strongylura leiura (Bleeker, 1850) Suezichthys bifurcatus Russell, 1986 Taratretis derwentensis Last, 1978 Thalasseleotris adela Hoese & Larson, 1987 Thamnaconus degeni (Regan, 1903) Thunnus alalunga (Bonnaterre, 1788) Thunnus maccoyii Castelnau, 1872) Thunnus obesus (Lowe, 1839) Thyrsites atun (Euphrasen, 1791) Thysanophrys cirronasus ( = Thysanophrys cirronasa) (Richardson, 1848) Tilodon sexfasciatum = Tilodon sexfasciatus (Richardson, 1842) Torquigener pleurogramma (Regan, 1903) Trachichthys australis Shaw & Nodder, 1799 163 Largescale Saury / Checkered Lizardfish Ladder Eel Woodward‘s Pomfret / Western Pomfred Rough Leatherjacket Short-finned Worm-eel / Shortfinned Worm Eel Short-headed Worm-eel / Long-finned Worm-eel / Longfinned Worm Eel Blue Mackerel King Gar Southern Red Scorpionfish Sea Sweep Banded Sweep Samson Fish Yellowtail Kingfish Blue Warehou Silver Warehou / Spotted Warehou King George Whiting Silver Whiting / Sand Whiting / Southern School Whiting / School Whiting Yellowfin Whiting / Yellow-finned Whiting Wood‘s Siphonfish Tubemouth Slender Weed Whiting / Short-nose Weed Whiting Pencil Weed Whiting / Long-nose Weed Whiting Sharp-nosed Weed Whiting / Sharpnose Weed Whiting Long-rayed Weed Whiting / Long-rayed Rock Whiting / Longray Rock Whiting Long-tailed Weed Whiting / Long-tail Weed Whiting / Longtail Weed Whiting Snook / Short-finned Seapike / Shortfin Seapike Blue Sprat Dusky Crawler Sand Crawler Spotted Pipefish Wide-bodied Pipefish / Wide-body Pipefish / Widebody Pipefish Slender Longtom / Banded Needlefish Striped Trawl Wrasse Derwent Flounder Cryptic Sea Gudgeon / Dusky Marine Gudgeon / Marine Gudgeon Degen‘s Leatherjacket / Blue-finned Leatherjacket Albacore Southern Bluefin Tuna Bigeye Tuna Barracouta Rock Flathead / Tassel-snouted Flathead Moonlighter / Six-banded Coralfish / Six-banded Coral Fish Weeping Toado / Banded Toadfish / Common Blowfish Roughy / Southern Roughy Trachinops noarlungae Glover, 1974 Trachurus declivis (Jenyns, 1841) Trachurus novaezelandiae Richardson, 1843 Trianectes bucephalus McCulloch & Waite, 1918 Trinorfolkia clarkei = Norfolkia clarkei (Moreton, 1888) Trinorfolkia cristata = Norfolkia cristata (Kuiter, 1986) E Trinorfolkia incisa = Norfolkia incisa (Kuiter, 1986) Upeneichthys vlamingii (Cuvier, 1829) Urocampus carinirostris Castelnau, 1872 P Vanacampus margaritifer (Peters, 1868) P Vanacampus phillipi (Lucas, 1891) P Vanacampus poecilolaemus (Peters, 1868) P Vincentia badia Allen, 1987 Vincentia conspersa (Klunzinger, 1872) Vincentia macrocauda Allen, 1987 Zanclistius elevatus (Ramsay & Ogilby, 1888) Zebrias penescalaris Gomon, 1987 Zeus faber Linnaeus, 1758 164 Yellow-headed Hulafish / Noarlunga Hulafish Jack Mackerel / Horse Mackerel / Cowanyoung Yellowtail Scad / Yellowtail Horse Mackerel Bighead Triplefin / Bighead Threefin / Bullhead Triplefin / Bullhead Threefin Common Triplefin / Common Threefin /Clarke‘s Triplefin / Clarke‘s Threefin Crested Threefin / Crested Triplefin Notched Threefin / Notched Triplefin Red Mullet / Blue-spotted Goatfish / Southern Goatfish Hairy Pipefish Mother-of-Pearl Pipefish Port Phillip Pipefish Long-Snout Pipefish Scarlet Cardinalfish Southern Cardinalfish Smooth Cardinalfish Blackspot Boarfish / Black-spotted Boarfish / Longfin Boarfish Duskybanded Sole / Dusky-banded Sole John Dory Appendix 6: Cartilaginous fish species likely to occur in the marine component of the AW NRM region (to 3 nautical miles from shore), with notes on conservation status. V = oceanic species; rare inshore, as vagrant. E = apparently endemic within South Australia. From Daley et al. (2006); Last et al. (2008); Gomon et al., (2008); CSIRO (2009); ABRS (2009). International: IUCN = World Conservation Union Red List (2009 version): CR = Critically Endangered; En = Endangered; Vul = Vulnerable; LR/NT = Lower Risk but Near Threatened (revised as NT = Near Threatened, from 2002 onwards); LC = Least Concern; DD = Data Deficient. National: EPBC = Commonwealth Environment Protection and Biodiversity Conservation Act 1999: Cr = Critically Endangered; En = Endangered; Vul = Vulnerable; Con = Conservation Dependent; Mar = Listed Marine Species, under s248 of the EPBC Act 1999; Mig = Listed Migratory Species under EPBC Act 1999. State: P = Protected under the Fisheries Management (General) Regulations 2007 of the Fisheries Management Act 2007 in South Australia. Scientific Name Alopias vulpinus Common Name Thresher Shark Asymbolus occiduus Southern Shovelnose Ray / Western Shovelnose Ray Western Spotted Catshark Asymbolus vincenti Gulf Catshark Callorhinchus milii Elephant Fish / Elephantfish / Elephant Shark Carcharhinus brachyurus Bronze Whaler Carcharhinus obscurus Dusky Shark / Black Whaler Carcharias taurus Grey Nurse Shark Aptychotrema vincentiana Carcharodon carcharias P White Shark Conservation Status IUCN-Vul (A2bd+3bd+4bd) IUCN-LC IUCN-LC IUCN-LC IUCN-LC IUCN-NT IUCN-Vul (A2bd) IUCN-Vul (A2cd+3cd) EPBC-Vul, EPBC-Mig IUCN-LC IUCN-Vul (A2ad+3d) IUCN-LC IUCN-DD Cephaloscyllium laticeps Draughtboard Shark Cetorhinus maximus V Basking Shark Dasyatis brevicaudata Smooth Stingray Dasyatis thetidis Black Stingray Dentiraja flindersi Dipturus cerva Pygmy Thornback Skate White-spotted Skate - Dipturus confusus Dipturus oculus Dipturus whitleyi Long-nose Skate Oscellate Skate Melbourne Skate - Galeorhinus galeus School Shark Galeus boardmani / Figaro boardmani Furgaleus macki Sawtail Shark Heptranchias perlo Sharpnose Sevengill Shark Heterodontus portusjacksoni Port Jackson Shark Hypnos monopterygium Coffin Ray Hypogaleus hyugaensis Pencil Shark Irolita waitii Southern Round Skate Isurus oxyrinchus Shortfin Mako Lamna nasus Porbeagle Narcine tasmaniensis Tasmanian Numbfish Notorynchus cepedianus Broadnose Sevengill Shark IUCN-NT IUCN-Vul (A2bd+4bd) IUCN-Vul (A2bd+3d+4bd) EPBC-Con IUCN-LC Whiskery Shark 165 IUCN-LC IUCN-NT IUCN-LC IUCN-LC IUCN-NT IUCN-LC IUCN-VU (A2bd+4d) Indowest Pacific subpopulation IUCN-VU (A2bd+3d+4bd) IUCN-LC IUCN-DD Orectolobus maculatus Southern Banded Wobbegong / Large Ornate Wobbegong / Banded Wobbegong Spotted Wobbegong Mustelus antarcticus Gummy Shark Myliobatis australis Southern Eagle Ray Parascyllium ferrugineum Parascyllium variolatum Pavoraja nitida Prionace glauca Pristiophorus cirratus Pristiophorus nudipinnis Pteroplatytrygon violacea = Dasyatis guileri Sphyrna zygaena Squalus acanthias Rusty Carpetshark Varied Carpetshark Peacock Skate Blue Shark Common Sawshark Southern Sawshark Pelagic Stingray Squalus chloroculus Squalus megalops Greeneye Dogfish / Greeneye Spurdog Piked Spurdog / Piked Dogfish / Shortnose Spurdog Australian Angelshark Ornate Angelshark / Ornate Angel Shark Cobbler Wobbegong Short-tail Torpedo Ray Western Shovelnose Stingaree Bight Stingaree Southern Fiddler Ray Spotted Stingaree Coastal Stingaree Sparsely-spotted Stingaree Orectolobus halei Squatina australis Squatina tergocellata Sutorectus tentaculatus Torpedo macneilli Trygonoptera mucosa Trygonoptera ovalis Trygonorrhina fasciata Urolophus gigas Urolophus orarius E Urolophus paucimaculatus Smooth Hammerhead White-spotted Dogfish / White-spotted Spurdog 166 IUCN-NT IUCN-NT IUCN-LC IUCN-LC IUCN-LC IUCN-LC IUCN-NT IUCN-NT IUCN-LC IUCN-LC IUCN-LC IUCN-VU (A2bd+3bd+4bd) IUCN-LC (Australasia subpopulation) IUCN-NT IUCN-DD IUCN-LC IUCN-LC IUCN-LC IUCN-DD IUCN-LC IUCN-LC IUCN-LC IUCN-LC IUCN-EN (B1ab(v)) IUCN-LC Appendix 7: Marine mammals reported to occur in the Great Australian Bight, including Head of the Bight area (from Ling, 1991; Kemper and Ling, 1991; Kemper et al., 2005; Director of National Parks, Commonwealth DEH, 2005; DEWHA, 2010). A = recorded strandings east of AW NRM region (e.g. Fowlers Bay or Ceduna area). Notes on conservation status are as follows: International: IUCN = World Conservation Union Red List (2009 version): CR = Critically Endangered; En = Endangered; Vul = Vulnerable; LR/NT = Lower Risk but Near Threatened (revised as NT = Near Threatened, from 2002 onwards); LC = Least Concern; DD = Data Deficient. National: EPBC = Commonwealth Environment Protection and Biodiversity Conservation Act 1999: Cr = Critically Endangered; En = Endangered; Vul = Vulnerable; Mar = Listed Marine Species, under s248 of the EPBC Act 1999; Mig = Listed Migratory Species under EPBC Act 1999. South Australia: NPW = South Australian National Parks and Wildlife Act 1972: En = Endangered (Schedule 7); Vul = Vulnerable (Schedule 8); Rare = (Schedule 9). Latin Name Common Name Conservation Status New Zealand Fur Seal Arctocephalus forsteri IUCN-LC Minke Whale A Balaenoptera acutorostrata IUCN-LC; NPW-Rare Blue Whale Balaenoptera musculus IUCN-Endg (A1abd); EPBC-Endg; EPBC-Mig; NPW-Endg Pygmy Blue Whale Balaenoptera musculus brevicauda IUCN-Endg (A1abd) (as Blue Whale B. musculus); EPBC-Endg (as Blue Whale B. musculus) NPW-Endg (as Blue Whale B. musculus) Fin Whale Balaenoptera physalus IUCN-Endg (A1d); EPBC-Vul; EPBCMig; NPW-Vul Common Short-beaked Dolphin Delphinus delphis IUCN-LC Southern Right Whale Eubalaena australis IUCN-LC; EPBC-Endg; EPBC-Mig; NPW-Vul Short-finned Pilot Whale Globicephala macrorhynchus IUCN-DD; NPW-Rare Long-finned Pilot Whale A Globicephala melas IUCN-DD Risso‘s Dolphin Grampus griseus IUCN-LC; NPW-Rare Southern Bottlenose Whale A Hyperoodon planifrons IUCN-LC; NPW-Rare Pygmy Sperm Whale Kogia breviceps IUCN-DD; NPW-Rare 167 Humpback Whale A Megaptera novaeangliae IUCN-LC globally; IUCN-Endg (A1ad) (Oceania sub-population); EPBC-Vul; EPBC-Mig; NPW-Vul Gray‘s Beaked Whale Mesoplodon grayi IUCN-DD; NPW-Rare Straptooth Whale Mesoplodon layardii IUCN-DD; (an unidentified beaked whale) Mesoplodon sp. Australian Sea Lion Neophoca cinerea IUCN-Endg (A2bd+3d);EPBC-Vul; EPBC-Mar; NPW-Vul New Zealand Fur Seal Arctocephalus forsteri IUCN-LC; EPBC-Mar; Killer Whale Orcinus orca IUCN-DD; EPBC-Mig Sperm Whale Physeter macrocephalus IUCN-Vul (A1d); EPBC-Mig; NPWRare False Killer Whale Pseudorca crassidens IUCN-DD; NPW-Rare Indo-Pacific Bottlenose Dolphin Tursiops aduncus IUCN-DD; Common Bottlenose Dolphin Tursiops truncatus IUCN-LC 168 Appendix 8: Species of marine and/or coastal bird, occurring in the AW NRM area (from Director of National Parks, Commonwealth DEH, 2005; Caton et al., 2007; Dennis, 2008; AG DEWHA, 2009), with notes on status. U = unverified; N = not confined to coastal area, and also occurs inland. International: IUCN = World Conservation Union Red List (2009 version): CR = Critically Endangered; En = Endangered; Vul = Vulnerable; LR/NT = Lower Risk but Near Threatened (revised as NT = Near Threatened, from 2002 onwards); LC = Least Concern; DD = Data Deficient. JAMBA = Japan-Australia Migratory Bird Agreement CAMBA = China-Australia Migratory Bird Agreement ROKAMBA = Republic of Korea-Australia Migratory Bird Agreement National: EPBC = Commonwealth Environment Protection and Biodiversity Conservation Act 1999: Cr = Critically Endangered; En = Endangered; Vul = Vulnerable; Mar = Listed Marine Species, under s248 of the EPBC Act 1999; Mig = Listed Migratory Species under EPBC Act 1999. South Australia: NPW = South Australian National Parks and Wildlife Act 1972: En = Endangered (Schedule 7); Vul = Vulnerable (Schedule 8); Rare = (Schedule 9). Latin Name Common Name Conservation Status Actitis hypoleucos (=Tringa hypoleucos) Common Sandpiper IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mig; EPBC-Mar; NPW-Rare Ardenna grisea Sooty Shearwater IUCN-NT; EPBC-Mig; EPBC-Mar; CAMBA; JAMBA (= Puffinus carneipes) Flesh-footed Shearwater / Fleshy-footed Shearwater IUCN-LC; EPBC-Mig; EPBC-Mar; JAMBA; ROKAMBA; NPW-Rare Ardenna tenuirostris Short-tailed Shearwater IUCN-LC; EPBC-Mig; EPBC-Mar; JAMBA; ROKAMBA Arenaria interpres Ruddy Turnstone IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mar; NPW-Rare Calidris acuminate Sharp-tailed Sandpiper IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mig; EPBC-Mar Calidris alba Sanderling Calidris canutus Red Knot IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mar Calidris ruficollis Red-necked Stint IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mig; EPBC-Mar Calidris tenuirostris Great Knot IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mig; EPBC-Mar; NPW-Rare Catharacta skua Great Skua IUCN-LC; EPBC-Mar Charadrius ruficapillus Red-capped Plover IUCN-LC; EPBC-Mar Cladorhynchus leucocephalus Banded Stilt IUCN-LC; NPW-Vul (= Puffinus griseus) Ardenna carneipes (= Puffinus tenuirostris) 169 Diomedea epomophora epomophora Southern Royal Albatross IUCN-VU (D2); EPBC-Vul; EPBCMig; EPBC-Mar; NPW-Vul Diomedea exulans Wandering Albatross IUCN-VU (A4bd) (with decreasing population trend) JAMBA; EPBCVul; EPBC-Mig; EPBC-Mar; NPWVul Egretta sacra Eastern Reef Egret IUCN-LC; CAMBA; EPBC-Mig; EPBC-Mar; NPW-Rare Erythrogonys cinctus Red-kneed Dotterel IUCN-LC Eudyptula minor Little Penguin IUCN-LC; EPBC-Mar Falco peregrines N Peregrine Falcon IUCN-LC; EPBC-Mig; NPW-Rare Haematopus longirostris Australian Pied Oystercatcher IUCN-LC; NPW-Rare Haematopus fuliginosus Sooty Oystercatcher IUCN-LC; NPW-Rare Haliaeetus leucogaster White-bellied Sea-Eagle IUCN-LC; CAMBA; EPBC-Mig; EPBC-Mar; NPW-En Heteroscelus brevipes / Tringa incana / Tringa brevipes Grey-tailed Tattler IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mig; EPBC-Mar; NPW-Rare Larus dominicanus U Kelp Gull IUCN-LC; EPBC-Mar; NPW-Rare Larus novaehollandiae Silver Gull IUCN-LC; EPBC-Mar Larus pacificus Pacific Gull IUCN-LC; EPBC-Mar Macronectes giganteus Southern Giant-Petrel IUCN-LC (but with decreasing population trend); EPBC-En; EPBC-Mar; NPW-Vul Morus serrator Australasian Gannet IUCN-LC; EPBC-Mar Numenius phaeopus Whimbrel IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mig; EPBC-Mar; NPW-Rare Platalea regia Royal Spoonbill IUCN-LC Pandion haliaetus Osprey / Eastern Osprey IUCN-LC; EPBC-Mig; EPBC-Mar; NPW-En Phoebetria fusca Sooty Albatross IUCN-EN (A4bd); EPBC-Vul; EPBC-Mig; EPBC-Mar; NPW-En Pluvialis squatarola Grey Plover IUCN-LC; JAMBA; CAMBA; ROKAMBA; EPBC-Mar; EPBC-Mig Procellaria aequinoctialis White-chinned Petrel IUCN-VU (A4bcde); EPBC-Mar; EPBC-Mig (Bonn) Recurvirostra novaehollandiae Red-necked Avocet IUCN-LC; EPBC-Mig; EPBC-Mar Sterna caspia Caspian Tern IUCN-LC; CAMBA; EPBC-Mig; EPBC-Mar Thalassarche cauta cauta Shy Albatross / Tasmanian Shy Albatross IUCN-NT; EPBC-Vul; EPBC-Mig; EPBC-Mar; NPW-Vul 170 Thalassarche chlororhynchos / Diomedea chlororhynchos Yellow-nosed Albatross / Atlantic Yellow-nosed Albatross IUCN-EN (A4bd; B2ab(v)) (with decreasing population trend); EPBC-Mig; EPBC-Mar; NPW-En Thalassarche melanophrys Black-browed Albatross IUCN-EN (A4bd) (with decreasing population trend); EPBC-Vul; EPBC-Mig; EPBC-Mar; NPW-Vul (as Campbell Island sub-species: Diomedea melanophrys impavida) Thalasseus bergii (= Sterna bergii) Crested Tern IUCN-LC; JAMBA; EPBC-Mar Thinornis rubricollis rubricollis Hooded Plover (eastern sub-species) IUCN-NT (for whole species, with decreasing population trend); EPBC-Mar; NPW-Vul 171