Water resource assessment for the Roper catchment Australia’s National Science Agency A report from the CSIRO Roper River Water Resource Assessment for the National Water Grid Editors: Ian Watson, Cuan Petheram, Caroline Bruce and Chris Chilcott ISBN 978-1-4863-1905-3 (print) ISBN 978-1-4863-1906-0 (online) Citation Watson I, Petheram C, Bruce C and Chilcott C (eds) (2023) Water resource assessment for the Roper catchment. A report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Chapters should be cited in the format of the following example: Petheram C, Bruce C and Watson I (2023) Chapter 1: Preamble: The Roper River Water Resource Assessment. In: Watson I, Petheram C, Bruce C and Chilcott C (eds) (2023) Water resource assessment for the Roper catchment. A report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Copyright © Commonwealth Scientific and Industrial Research Organisation 2023. To the extent permitted by law, all rights are reserved and no part of this publication covered by copyright may be reproduced or copied in any form or by any means except with the written permission of CSIRO. Important disclaimer CSIRO advises that the information contained in this publication comprises general statements based on scientific research. The reader is advised and needs to be aware that such information may be incomplete or unable to be used in any specific situation. No reliance or actions must therefore be made on that information without seeking prior expert professional, scientific and technical advice. To the extent permitted by law, CSIRO (including its employees and consultants) excludes all liability to any person for any consequences, including but not limited to all losses, damages, costs, expenses and any other compensation, arising directly or indirectly from using this publication (in part or in whole) and any information or material contained in it. CSIRO is committed to providing web accessible content wherever possible. If you are having difficulties with accessing this document, please contact Email CSIRO Enquiries . CSIRO Roper River Water Resource Assessment acknowledgements This report was funded through the National Water Grid’s Science Program, which sits within the Australian Government’s Department of Climate Change, Energy, the Environment and Water. Aspects of the Assessment have been undertaken in conjunction with the Northern Territory Government. The Assessment was guided by two committees: i. The Assessment’s Governance Committee: CRC for Northern Australia/James Cook University; CSIRO; National Water Grid (Department of Climate Change, Energy, the Environment and Water); NT Department of Environment, Parks and Water Security; NT Department of Industry, Tourism and Trade; Office of Northern Australia; Qld Department of Agriculture and Fisheries; Qld Department of Regional Development, Manufacturing and Water ii. The Assessment’s joint Roper and Victoria River catchments Steering Committee: Amateur Fishermen’s Association of the NT; Austrade; Centrefarm; CSIRO, National Water Grid (Department of Climate Change, Energy, the Environment and Water); Northern Land Council; NT Cattlemen’s Association; NT Department of Environment, Parks Australia; Parks and Water Security; NT Department of Industry, Tourism and Trade; Regional Development Australia; NT Farmers; NT Seafood Council; Office of Northern Australia; Roper Gulf Regional Council Shire Responsibility for the Assessment’s content lies with CSIRO. The Assessment’s committees did not have an opportunity to review the Assessment results or outputs prior to its release. This report was reviewed by Kevin Devlin (Independent consultant). For further acknowledgements, see page xxii. Acknowledgement of Country CSIRO acknowledges the Traditional Owners of the lands, seas and waters of the area that we live and work on across Australia. We acknowledge their continuing connection to their culture and pay our respects to their Elders past and present. Photo Looking along the Roper River at Red Rock, Northern Territory. Source: CSIRO – Nathan Dyer 3 Living and built environment of the Roper catchment Authors: Pethie Lyons, Danial Stratford, Chris Stokes, Diane Jarvis, Rob Kenyon, Jodie Pritchard, Linda Merrin, Simon Linke, Rocio Ponce Reyes, Caroline Bruce, Heather McGinness and Andrew Taylor Chapter 3 discusses a wide range of considerations relating to the living component of the catchment of the Roper River and the environments that support these components; the people who live in the catchment or have strong ties to it and the existing transport, power and water infrastructure. The key components and concepts of Chapter 3 are shown in Figure 3-1. Figure 3-1 Schematic diagram of key components of the living and built environment to be considered in the establishment of a greenfield irrigation development For more information on this figure or equation, please contact CSIRO on enquiries@csiro.au Numbers refer to sections in this chapter. 3.1 Summary This chapter provides information on the living and built environment including information about the people, the ecology, the infrastructure and the institutional context of the Roper catchment. It also examines the values, rights, interests and development objectives of Indigenous people. 3.1.1 Key findings Ecology The largely intact habitats and landscapes of the Roper catchment provide near-natural ecosystem services that support high biodiversity, recreational activities, tourism, traditional and commercial fisheries, and areas of agricultural production. Within the freshwater sections of the Roper catchment are extensive areas with high habitat values including ephemeral and persistent rivers, wetlands, floodplains and groundwater-dependent ecosystems (GDEs), including the Directory of Important Wetlands in Australia (DIWA) listed Mataranka Thermal Pools. For the marine and estuarine environments, the Roper River provides some of the largest flows into the western Gulf of Carpentaria, supporting extensive intertidal, estuarine and marine communities including those in the Limmen Bight Marine Park. Flows from the Roper River into the Gulf of Carpentaria support recreational and commercial fisheries including barramundi (Lates calcarifer) and the northern common banana prawn fishery (Penaeus merguiensis and P. indicus). The habitats of the Roper catchment contain some of northern Australia’s most iconic wildlife species, including barramundi, freshwater sawfish (Pristis pristis) and dugong (Dugong dugong), as well as many lesser known plants and animals that are also of great conservation significance. Among the diversity in the Roper catchment are observations of over 130 species of fish, and salt flats, wetlands and floodplains providing habitat for often tens of thousands of waterbirds. Changes in land and water resources can have serious consequences for the ecology of rivers. Water resource development that results in changes to the magnitude, timing and duration of both low and high flows can affect species, habitats and ecological processes such as connectivity. Water resource development can also facilitate or exacerbate other impacts, including the spread or establishment of invasive species, increases in other pressures, and changes to water quality, including the availability and distribution of nutrients. Demographics, industries and infrastructure The Roper catchment has a population of about 2500, with a population density 100 times lower than that of Australia as a whole. The region contains no large urban centres, however, there are several small towns and communities within the catchment including Barunga, Beswick, Bulman, Daly Waters, Larrimah, Mataranka (the regional centre), Minyerri and Ngukurr. The only one of these settlements with a population greater than 1000 is Ngukurr (population about 1100). The typical resident of the region is younger, poorer and more likely to identify as Indigenous than the typical resident of the NT and of Australia as a whole. The main land uses across the catchment are for conservation (49%) and grazing (46%), noting that in terms of tenure, 45% of the catchment is held as Aboriginal freehold. The gross value of agricultural production (GVAP) in the Roper catchment is approximately $73 million. Beef cattle contribute around $55 million to GVAP and cropping accounts for the remaining $18 million. The Roper catchment is characterised by a sparse network of major roads. The Stuart Highway is the most important road, connecting to Darwin in the north and Adelaide to the south. All roads within the Roper catchment permit Type 2 road trains (vehicles up to 53 m in length) that then have onward access, via northern routes, to Darwin Port. There is a good quality standard gauge rail line through the west of the catchment that provides freight access to the port. The Darwin- Katherine Interconnected System (DKIS) electricity transmission network reaches the western edge of the Roper catchment, passing through Mataranka and reaching as far south as Larrimah. A small branch off this main transmission line serves Barunga (Bamyili) and Beswick, and a distribution line links Jilkminggan to nearby Mataranka. Most of the Roper catchment, however, is too remote to be covered by the DKIS. The largest three off-grid remote communities rely on hybrid electricity systems powered by diesel generators supplemented with solar: Ngukurr (400 kW solar system), Minyerri (275 kW) and Bulman (100 kW). There are no major dams or water transmission pipelines in the Roper catchment. Urban water for domestic consumption therefore depends mainly on treated groundwater (from bores) as the preferred source for larger settlements. Indigenous values and development objectives This activity addresses the existing information needs with respect to Indigenous water issues in the Assessment area to provide foundations for further community and government planning and decision making. This activity provides a regionally specific assessment designed to help non-Indigenous decision makers understand general Indigenous valuations of water, wider connections to country, and the rights and interests attached to those. It highlights likely issues to be raised in future discussions with Indigenous groups about development proposals, community planning and Indigenous business objectives. This activity highlighted key conceptual issues and principles with respect to Indigenous people and generated a representative set of Indigenous water values, rights and interests. It focused on data gathering and individual consultations. It did not attempt to conduct community-based planning or to identify formal Indigenous group positions on any of the matters raised. The research approach was primarily based on face-to-face interviews with senior members of the Wubalawan, Mangarrayi, Bagala, Dalabon, Ngalakan, Ngandi, Warndarrang and Alawa groups. The research describes some key concepts and principles as they relate to Indigenous Australians. These include Indigenous perspectives on engagement, ‘culture’, ‘values, rights and interests’ and understandings of ‘development’. Indigenous people and the groups they belong to have significant land holdings and rights in country through the Aboriginal Land Rights (Northern Territory) Act 1976 (Cth) (ALRA), the Northern Territory Aboriginal Sacred Sites Act 1989 and native title determinations. These holdings are an important focus for discussions about water and about sustainable development in the Roper catchment (also see Macintosh et al. (2018) for a legal, policy and regulatory analysis of water development in northern Australia). Indigenous objectives combine economic viability and sustainability with a range of wider social, cultural and environmental goals. Participants in the activity provided crucial framing information about Indigenous culture, country and people. Particular obligations to past and future generations to maintain customary practices and knowledge and care for the country properly are identified. These obligations entail responsibilities to near neighbours and downstream groups. The overall importance of water is demonstrated by clear statements from the research participants. Key water issues for Indigenous people in the Roper catchment include: • ensuring enough water and of sufficient quality to maintain healthy landscapes (environmental flows) and sustain cultural resources and practices • monitoring and reporting of water uses, availability and development impacts on water quality for informed decision making about future development • maintaining adequate and good quality supplies for human consumption and recreation in communities, for outstations and to maintain green shaded community spaces • securing sufficient water reserves for current and future economic activity. In 2019, the Northern Territory Government introduced the Strategic Aboriginal Water Reserves policy under the Water Act 1992 (NT) to improve access to water allocations for Aboriginal people holding land under the ALRA. However, not all Aboriginal people in the Roper catchment have such rights. There remains relatively limited means for Indigenous knowledge of water to be expressed in public policy and planning. Indigenous peoples’ knowledge of formal government-led water planning in the area was found to be relatively low. Cultural heritage impacts from development are a significant issue. Results from the Indigenous participants in the activity showed that, if water development were to occur, the general trend from most favourable to least favourable forms of development would be: flood harvesting into smaller, offstream storages; sustainable bore and groundwater extraction; smaller instream dams inside tributaries or ancillary branches; and large instream dams in the river channels. With respect to Indigenous objectives and development planning, several interrelated development goals are identified and include management and control over water and improvements in the overall social and economic status of Indigenous people. There are clear relationships between access to secure clean water for community, community wellbeing and health, and development possibilities. In relation to wider development, group or community-based planning can help communities prioritise options for development. These can include establishing stand-alone Indigenous businesses and training outcomes such as local and regional resource monitoring and reporting programs. Indigenous people in the Roper catchment possess valuable natural and cultural assets and represent a significant potential labour force, but collectively lack business development skills and expertise. Indigenous development objectives, and Indigenous development partnerships, are best progressed through locally specific, group and community-based planning and prioritisation processes that are nested in a system of regional coordination. Indigenous people can also act as a substantial enabler of appropriate development. They seek to be engaged early and continuously in defining development pathways and options. 3.1.2 Introduction This chapter seeks to address the question ‘What are the existing: ecological systems; demographic and economic profile, land use, industries and infrastructure; and the values, rights, interests and development objectives of Indigenous people in the Roper catchment?’ The chapter is structured as follows: • Section 3.2 examines the ecological systems and assets of the Roper catchment, including the key habitats and key biota, and their important interactions and connections. • Section 3.3 examines the socio-economic profile of the Roper catchment including the current demographics, existing industries and infrastructure of relevance to water resource development. • Section 3.4 examines the Indigenous values, rights, interests and development objectives of Traditional Owners from the Roper catchment, generated through direct participation in the Assessment. 3.2 Roper catchment and its environmental values This section provides an overview of the environmental values, and freshwater, marine and terrestrial ecological assets in the Roper catchment. Unless otherwise stated, the material in this section is based on findings described in the companion technical report on ecological assets (Stratford et al., 2022). The comparatively intact landscapes of the Roper catchment are important for the ecosystem services they provide, including recreational activities, tourism, traditional and commercial fisheries, and areas of agricultural production, notably cattle grazing on native pastures. In addition, they hold important ecological and environmental values. The Roper River is a large perennial river and drains an area of 77,400 km2, one of the largest catchment areas flowing into the western Gulf of Carpentaria. Within this catchment and the surrounding marine environment are rich and important ecological assets including species, ecological communities, habitats and ecological processes and functions (Figure 3-2 presents a conceptualised summary of ecological values and assets found in the Roper catchment). The ecology of the Roper catchment is maintained by the river’s flow regime, shaped by the region’s wet-dry climate and the catchment’s complex geomorphology and topography, and driven by seasonal rainfall, evapotranspiration and groundwater discharge. Figure 3-2 Conceptual diagram of selected ecological values and assets of the Roper catchment Ecological assets include species of significance, species groups, important habitats and ecological process. See Table 3-1 for a complete list of the freshwater-dependent, marine and terrestrial ecological assets considered in the Roper catchment. Biota icons: Integration and Applicaiton Network (2022) Much of the natural environment of the Roper catchment is low relief, consisting of open woodlands, with escarpments, gorges and plateaux occurring across parts of the catchment. The wet-dry tropical climate results in highly seasonal river flow with 96% of rainfall between 1 November and 30 April (Section 2.4). The dynamic occurring between wet and dry seasons provides both challenges and opportunities for biota (Warfe et al., 2011). During the dry season, river flows are reduced and the streams in the catchment recede, many to isolated pools. However, in parts of the Roper catchment the persistence of water during the dry season is supported by discharge from aquifers including the Tindall Limestone Aquifer and the Dook Creek Formation (Faulks, 2001). In the dry season, the streams and waterholes (Figure 3-3) that persist, including the important spring-fed streams between Mataranka Thermal Pools and Red Lily Lagoon, provide critical refuge habitat for many aquatic species (Barber and Jackson, 2012; Faulks, 2001). In this respect, the Roper catchment is atypical of many of the other catchments of the wet- dry tropics (Kennard, 2010; Pettit et al., 2017) in having many tributaries supplemented by groundwater discharges (Faulks, 2001; Pettit et al., 2017). Due to the flat topography in parts of the catchment, the Roper River contains sections that braid into smaller channels. These braids and anabranches provide a diverse habitat structure. During the wet season, flooding inundates significant parts of the catchment connecting wetlands to the river channel, inundating floodplains and driving a productivity boom. This flooding is particularly evident in the lower parts of the catchment, including the floodplains, wetlands and intertidal flats of the Limmen Bight (an inlet extending for 135 km between Groote Eylandt and the Sir Edward Pellow Group), and delivers extensive discharges into the marine waters of the western Gulf of Carpentaria. For more information on this figure please contact CSIRO on enquiries@csiro.au Protected, listed and significant areas of the Roper catchment The protected areas located in the Roper catchment include two national parks, an Indigenous Protected Area and other conservation parks (Figure 3-4). Of the national parks, Elsey National Park covers approximately 140 km² and Limmen National Park approximately 9300 km², although much of Limmen National Park extends beyond the Roper catchment (Department of Agriculture‚ Water and the Environment, 2020b). The South East Arnhem Land Indigenous Protected Area covers an area of approximately 18,000 km2 and also extends beyond the Roper catchment. Also within the Roper catchment are the Wongalara Sanctuary and the St Vidgeon management area (approximately 2000 km2 and 2800 km2, respectively) (Department of Agriculture‚ Water and the Environment, 2020b). In the Roper catchment marine region are two contiguous marine parks, Limmen Bight in Territory waters and the Limmen Marine Park in Commonwealth waters, covering an area of approximately 870 km2 and 1400 km2, respectively. Further out in the Gulf of Carpentaria is the 7300 km2 Anindilyakwa Indigenous Protected Area consisting of Groote Eylandt and the surrounding waters (Department of Agriculture‚ Water and the Environment, 2020a). Figure 3-3 Waterlily (Nymphaea violacea) common to northern Australia found in billabongs, waterholes and rivers Photo: CSIRO - Nathan Dyer Figure 3-4 Location of protected areas and important wetlands within the Roper catchment Assessment area Includes management areas protected mainly for conservation through management intervention as defined by the International Union for Conservation of Nature. Dataset: Department of Agriculture‚ Water and the Environment (2020a, 2020b); Department of the Environment and Energy (2010) For more information on this figure please contact CSIRO on enquiries@csiro.au The Roper catchment includes two DIWA sites (Figure 3-4), namely the groundwater-fed Mataranka Thermal Pools and the coastal Limmen Bight (Port Roper) Tidal Wetlands System (Environment Australia, 2001; SKM, 2009). These two DIWA wetlands demonstrate a striking contrast between persistent freshwater riparian habitat and marine, coastal and near-shore habitats and reinforce the diversity of aquatic habitats that can be found within the Roper catchment. The Roper catchment contains no Ramsar-listed sites. The Mataranka Thermal Pools DIWA site is located in 4 ha of Elsey National Park in the upper Roper catchment that is maintained by permanent thermal springs (Department of Agriculture, Water and the Environment, 2019b). The artificially modified pools containing sections of paved and cemented areas are fringed by palm forest and drain into the nearby Waterhouse River (SKM, 2009). The pools provide stable habitat for flyspecked hardyhead (Craterocephalus stercusmuscarum) and chequered rainbowfish (Melanotaenia splendida inornata), while up to 200,000 little red flying-fox (Pteropus scapulatus) roost in the surrounding forest (Department of Agriculture, Water and the Envrionment, 2019b). Groundwater-dependent vegetation fringe the pools that are supported by persistent discharges from deep aquifers. The Limmen Bight Tidal Wetlands System DIWA site is part Aboriginal freehold and part private lease and is the second-largest area of saline coastal flats in the NT (1848 km2, excluding subtidal seagrass areas) (Department of Agriculture, Water and the Envrironment, 2019a). Limmen Bight forms a highly important habitat system of tidal wetlands (intertidal mud flats, saline coastal flats and estuaries) and while the whole site is tidal, it receives large volumes of freshwater inflows from the contributing catchments. The DIWA site provides one of the most important habitat sites nationally for dugongs (Palmer and Smit, 2019), as well as being an important habitat for several species of marine turtles (Department of Agriculture, Water and the Environment, 2019a). The seagrass beds of Limmen Bight are a major breeding area for prawns and help support an important fishing industry (Department of Agriculture, Water and the Environment, 2019a; Palmer and Smit, 2019; Tomkovish and Weston, 2008). Limmen Bight is also an important feeding ground for migratory shorebirds in the NT, with counts in the tens of thousands (Palmer and Smit, 2019). Shorebirds such as the eastern curlew (Numenius madagascariensis; Critically endangered) and great knot (Calidris tenuirostris; Critically endangered) migrate from their breeding grounds in the northern hemisphere and use the intertidal flats for feeding (Department of Agriculture, Water and the Environment, 2019a; Palmer and Smit, 2019; Tomkovish and Weston, 2008). Due to the provision of important habitat, Limmen Bight is a declared Important Bird Area by BirdLife International (BirdLife International, 2022). Important habitat types and values of the Roper catchment Within the freshwater sections of the Roper catchment are diverse habitats including persistent and ephemeral rivers, anabranches and braiding channels, wetlands, floodplains and GDEs (Faulks, 2001). The diversity and complexity of, and connection between, habitats within a catchment are vital for providing a range of habitat needs to support both aquatic and terrestrial biota (Schofield et al., 2018). In the wet season, flooding connects rivers to floodplains. Floodplain habitats, due to their water exchange during floods, support higher levels of primary and secondary productivity in comparison to surrounding areas with less frequent inundation (Pettit et al., 2011). Infiltration of water into the soil during the wet season and along persistent streams routinely enables riparian habitats to form an important interface between the aquatic and terrestrial environment. While riparian habitats often occupy a relatively small proportion of the catchment, they frequently have a higher abundance and species richness compared to surrounding habitats (Pettit et al., 2011; Xiang et al., 2016). The riparian habitats that fringe the rivers and streams of the Roper catchment are largely intact and include river red gum (Eucalyptus camaldulensis) overstory with Mataranka palm Livistona mariae rigida, Pandanus spp. and Melaleuca communities across many parts of the catchment (Faulks, 2001). Conversely, in the dry season, biodiversity is supported within the inchannel waterholes that persist in the landscape. Waterholes that remain become increasingly important as the dry season progresses and provide important refuge habitat for species and enable recolonisation into surrounding habitats upon the return of larger flows (Hermoso et al., 2013). Waterholes provide direct habitat for water-dependent species including fish, sawfish and turtles, as well as providing a source of water for other species more broadly within the landscape (McJannet et al., 2014; Waltham et al., 2013). GDEs occur across many parts of the Roper catchment and come in different forms, including aquatic, terrestrial and subterranean habitats. Aquatic GDEs, including DIWA-listed Mataranka Thermal Pools, contain springs and river sections that hold water throughout most dry seasons due to groundwater discharge. Aquatic GDEs are important for supporting aquatic life and fringing vegetation and in the wet-dry tropics often provide critical refuge during periods of the late dry season (James et al., 2013). Vegetation occurring adjacent to the waterways in the Roper catchment rely on water from a range of sources (surface water, soil water, groundwater) which are seasonally dynamic and highly spatially variable across floodplains. The sources of water may be from a combination of direct rainfall, bank recharge from instream flows, local floodplain recharge from surface water inundation during overbank flows, and/or shallow groundwater connected to intermediate or regional aquifer systems. Perennial floodplain vegetation often uses groundwater when it is within reach of the root network, particularly during the dry season or drought, but the origin of the groundwater used is only infrequently investigated (e.g. Canham et al., 2021). In most locations, vegetation is sustained by water available in unsaturated soils and never uses groundwater. However, in some locations, vegetation may use groundwater sourced from local alluvial recharge processes (including bank storage) or regional groundwater and this may be critical for maintaining vegetation condition. Sources of water used by vegetation can be patchy across floodplains (see the companion technical report on hydrogeological assessment (Taylor et al., 2023)) and vary from season to season. Subterranean aquatic ecosystems in the Roper catchment support a diverse and largely hidden stygofauna community within the Tindall Limestone karstic aquifer (Cambrian Limestone Aquifer; Oberprieler et al., 2021). Some subterranean species are distributed across a broad spatial range, while others have highly restricted ranges, which makes them more vulnerable to local changes where they occur (Oberprieler et al., 2021). For marine and estuarine environments, the Roper catchment, including the area of Limmen Bight and beyond, has extensive intertidal flats and estuarine communities, including mangroves, salt flats and seagrass habitats. These habitats are highly productive, have high cultural value and are often of national significance (Bradley, 2018; Department of Agriculture, Water and the Environment, 2019a; Poiner et al., 1987). The intertidal flats in the Roper catchment marine region are extensive, with the mangrove communities containing at least 19 woody plant species fringing the banks of streams and rivers (Palmer and Smit, 2019). Seagrass beds in nearby coastal Gulf of Carpentaria are of high diversity, are vigorous and provide an important food resource for dugongs, green turtles (Chelonia mydas) and prawns (Loneragan et al., 1997; Poiner et al., 1987). These near-coastal and estuary habitats support a major commercial barramundi fishery, while harvest of mud crabs (mainly Scylla serrata) also occurs along the coasts near Port Roper (Bayliss et al., 2014). Significant species and ecological communities of the Roper catchment The Roper catchment supports some of northern Australia’s most archetypical and important wildlife species, including sawfish (Vulnerable), marine turtles and dugong that occur in the coastal waters of the Gulf of Carpentaria. The regionally endemic Gulf snapping turtle (Elseya lavarackorum; Endangered) can be found associated with vegetated freshwater reaches of the catchment. Freshwater crocodiles (Crocodylus johnstoni) are common within the Roper River and its tributaries. While saltwater crocodiles (Crocodylus porosus) frequently occur in the lower Roper River to around Ngukurr, it can also be found upstream as far as near Elsey National Park (ALA, 2021). Across the catchment are many lesser known plants and animals that are also of great importance. Diversity in the Roper catchment is high, it is estimated to contain 270 vertebrate species (Dasgupta et al., 2019). Among the diversity in the Roper catchment are over 130 species of freshwater fishes, sharks and rays. Most of these fish species do not enter the marine environment and remain within the riverine and wetland habitats of the catchment. Owing to its healthy floodplain ecosystems and free-flowing rivers (Grill et al., 2019; Pettit et al., 2017), very few freshwater fishes in the study area are threatened with extinction. The Roper catchment is an important stopover habitat for migratory shorebird species that are listed under the Environment Protection and Biodiversity Conservation Act 1999 (EPBC Act) (Australian Government, 1999) including the northern Siberian bar-tailed godwit (Limosa lapponica menzbieri; Critically endangered), eastern curlew (Critically endangered) and the Australian painted snipe (Rostratula australis; Endangered). 3.2.1 Current condition and potential threats in the Roper catchment In conjunction with the diversity of landscapes, habitats, ecological communities and species of the Roper catchment there are also a range of economic enterprises, infrastructure and human impacts. The nature and extent to which human activities have modified the habitats and had an impact on species of the Roper catchment varies. Previous assessments have indicated the riverine habitat in the Roper catchment as being of high-quality condition and largely intact and unimpacted by clearing or development (Close et al., 2012; Faulks, 2001), although threatening processes continue to operate in more recent years, including impacts from pest species and as a result of other disturbances. Intertidal habitats including salt flats and mangroves are recognised as being of good condition and are often of national significance (Department of Agriculture, Water and the Environment, 2019a). Fishing in northern Australia is a valuable industry and the waters of the Gulf of Carpentaria contribute significantly to the national catch of important species, including banana prawns, mud crab and barramundi. The study area includes the towns of Ngukurr, Mataranka and Daly Waters, which provide Indigenous homelands, support a vital tourism industry and act as regional hubs for many of the properties across the catchment. While a moderate proportion of the catchment is under conservation reserves, the study area does face environmental threats. This includes the potential of tourism-related impacts at sensitive and vulnerable sites. In the Roper catchment the more significant or of higher concern impacts are largely localised and include areas such as Mataranka Thermal Pools (Department of Agriculture, Water and the Environment, 2019b). Northern Australia more broadly encompasses some of the last relatively undisturbed tropical riverine landscapes in the world, with low levels of flow regulation and low development intensity (Pettit et al., 2017; Vörösmarty et al., 2010). Riparian vegetation characteristics of the Roper catchment are considered to have not been affected by extensive clearing or development, although the impact that does occur is often associated with stock and pest species (Faulks, 2001). One of the most significant environmental threats to remote regions across northern Australia is that of introduced plants and animals. In the Roper catchment, cane toad (Rhinella marina), water buffalo (Bubalus bubalis) and wild pig (Sus scrofa) are among the introduced animals (ALA, 2021; Department of Agriculture, Water and the Environment, 2021a). Weeds of national significance in the aquatic systems of northern Australia include giant sensitive tree (Mimosa pigra), olive hymenachne (Hymenachne amplexicaulis), cabomba (Cabomba caroliniana), salvinia (Salvinia molesta) and rubber vine (Cryptostegia grandiflora) (Close et al., 2012). Weed species of interest in and around the Roper catchment include gamba grass (Andropogon gayanus), para grass (Brachiaria mutica), giant sensitive tree and prickly acacia (Vachellia nilotica) (Department of Agriculture, Water and the Environment, 2021a) and some of these, including sensitive tree and para grass, are recognised to have had a significant impact on undeveloped rivers more broadly in northern Australia (Davies et al., 2008). Water resource development and ecology Globally, water resource development has a range of known impacts on ecological systems. These impacts can involve flow regime change, longitudinal and lateral connectivity, habitat modification and loss, introduction and support of invasive and non-native species, and synergistic and co- occurring processes. Flow regime change Water resource development, including water harvesting and creating instream structures for water retention, can influence the timing, quality and quantity of water that is provided by catchment runoff into the river system. The natural flow regime including the magnitude, duration, timing, frequency and pattern of flow events is important in supporting a broad range of environmental processes upon which species and habitat condition depend (Lear et al., 2019; Poff et al., 1997). Flow conditions provide the physical habitat in streams and rivers, which determines biotic use and composition, to which life-history strategies are evolved, and which enables movement and migration between habitats and exchange of nutrients and materials (Bunn and Arthington, 2002; Jardine et al., 2015). In a river system, the natural periods of both low and high flow (including no-flow events) are important to support the natural function of habitats, their ecological processes and the shaping of biotic communities (King et al., 2015). Water resource development through the attenuation of flows can lead to impacts across significant distances downstream of the development, including into coastal and near-shore marine habitats (Broadley et al., 2020; Pollino et al., 2018). Longitudinal and lateral connectivity River flow facilitates the exchange of biota, materials, nutrients and carbon along the river and into the coastal areas (longitudinal connectivity), as well as between the river and the floodplain (lateral connectivity) (Pettit et al., 2017; Warfe et al., 2011). Physical barriers such as weirs and dams, or a reduction in the magnitude (and the duration or frequency) of flows can have an impact on longitudinal and lateral connectivity, changing the rate or timing of exchanges (Crook et al., 2015). These impacts can include changes in species’ migration and movement patterns, as well as altered erosion processes and discharges of nutrients into rivers and coastal waters (Brodie and Mitchell, 2005). Seasonal patterns and rates of connection and disconnection caused by flood pulses are important for providing seasonal habitat, enabling movement of biota into new habitats and their return to refuge habitats during drier conditions (Crook et al., 2019). Habitat modification and loss Water resource development can result in direct loss of habitat. This can include artificially creating lake habitat behind an impoundment, resulting in loss of terrestrial and stream habitat due to inundation by the impoundment. Agricultural development results in the conversion of habitat to more intensive agriculture. Infrastructure including roads and canals can lead to fragmentation of terrestrial habitat or the artificial connection of aquatic habitat that has been historically distinct. Invasive and non-native species Water resource development often results in homogenisation of flow or habitats. This can be due to the changed patterns of capture and release of flows or the creation of impoundments for storage and regulation. Invasive species are recognised to often be at an advantage in such modified habitats (Bunn and Arthington, 2002). Modified landscapes, such as lakes or the conversion of ephemeral streams into perennial streams, can be a pathway for introduction and support the establishment of non-native species (incidental, accidental or deliberate) including pest plant and fish species (Bunn and Arthington, 2002; Close et al., 2012; Ebner et al., 2020). Increased human activity can lead to increased risk of invasive species being introduced. Synergistic and co-occurring processes both local and global Along with water resource development comes a range of other pressures and threats, including increases in fishing, vehicles, habitat fragmentation, pesticides, fertilisers and other chemicals, erosion, degradation due to stock, changed fire regimes, climate change and other human disturbances both direct and indirect. Some of these pressures are the direct result of changes in land use associated with water resource development, others may occur regionally or globally and act synergistically with water resource development and agricultural development to increase the risk to species and their habitats (Craig et al., 2017; Pettit et al., 2012). To describe the ecology of the Roper catchment and discuss the likely impacts of future water resource development on this system, a suite of ecological assets has been selected (Table 3-1). Assets are classified as species, species groups or habitats and can be considered as either partially or fully freshwater-dependent, or terrestrial or marine dependent upon freshwater flows (or servicesprovidedby freshwater flows).This chapter considers a key subset of assets,as indicatedinTable3-1. More information on the ecological assetsofthe Roper catchmentand theirdistribution is available in the companion technical report on ecological assets(Stratford et al., 2022). In Chapter 7,resultsof the modelling andanalysis are presented toexplorethe potential ofchangetothese assets as a consequenceof water resource development. Table3-1Freshwater, marine and terrestrial ecological assets with freshwater dependences Anasterisk (*) represents an asset outlined in this report,with all listed species,speciesgroupsand habitatassetsdetailed in the companion technical report on ecological assets(Stratford et al., 2022). SYSTEMASSET TYPEFRESHWATERDEPENDENT ASSET FreshwaterImportant habitatFloodplain wetlands* Inchannel waterholes Species andspecies groupsFor more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au. CatfishGrunter* Freshwater turtles* Sawfish (also considered in marine) Waterbirds group:colonial and semi-colonial waders* Waterbirds group:cryptic wading waterbirdsWaterbirds group:shorebirds Waterbirds group:swimmers, divers and grazersMarineImportant habitatMangroves Saltpans andsalt flatsSeagrass Species andspecies groupsBanana prawns* Endeavour prawnsMud crabs* Tiger prawns Barramundi (also considered infreshwater)* Mullet Sawfish (also consideredin freshwater) ThreadfinTerrestrialImportant habitatGroundwater-dependent ecosystems Surface water dependent vegetation* 3.2.2Ecologicalassets fromfreshwater systems Thefreshwater systems in northern Australia containhigh diversity, with many unique andsignificant species andhabitats. The ecology of thefreshwater systems of the Roper catchmentis supported by, and adapted to,the highly seasonal flow regimes of the wet-dry tropics.Table3-1presentsthe full list of assetsevaluatedin the Roper catchment ecology activity, and this section 112|Water resource assessment for the Roper catchment provides information on a sample of these as relevant to the freshwater systems of this catchment. Floodplain wetlands Wetlands in the wet-dry tropics of Australia are considered to have great conservation value (Finlayson et al., 1999), and are considered one of the most diverse aquatic ecosystems in Australia (Douglas et al., 2005). Wetlands provide permanent, temporary or refugia habitat for both local and migratory waterbirds (van Dam et al., 2008) and spawning grounds and nurseries for floodplain-dependent fish (Ward and Stanford, 1995), as well as habitat for many other aquatic and riparian species (van Dam et al., 2008) (Figure 3-5). Floodplain wetlands are an important source of nutrients and organic carbon, driving primary and secondary productivity (Junk et al., 1989; Nielsen et al., 2015). Wetlands also provide a range of additional ecosystem services, including water quality improvement, carbon sequestration and flood mitigation (Mitsch et al., 2015). Hydrological regimes are fundamental to sustaining ecological characteristics of rivers and their associated floodplains (Pettit et al., 2017). In the wet-dry tropics of northern Australia, the ecology of wetlands is highly dependent on the seasonal rainfall-runoff pattern, and the associated low and high flows (Pidgeon and Humphrey, 1999; Warfe et al., 2011). These flows are important drivers of floodplain wetland ecosystem structure and processes (Close et al., 2012; Warfe et al., 2011). Changes to flow characteristics are likely to have a significant impact on the aquatic biota (Close et al., 2012). The timing, duration, extent and magnitude of wetland inundation has the greatest impact on the ecological values, including species diversity, productivity and habitat structure (Close et al., 2015). Under the Ramsar Convention a wetland is defined as (Ramsar Convention Secretariat, 2004): ‘areas of marsh, fen, peatland or water, whether natural or artificial, permanent or temporary, with water that is static or flowing, fresh, brackish or salt, including areas of marine water the depth of which at low tide does not exceed six metres.’ The Northern Territory Government defines wetlands as including coastal salt marshes, mangrove swamps, freshwater lakes and swamps, floodplains, freshwater ponds, springs and saline lakes that can be permanent, seasonal or intermittent, and can be natural or artificial (Northern Territory Government, 2020). For the purpose of this Assessment, we do not consider areas within the river channel as wetlands (considered as inchannel waterholes). Similarly, marine or saline habitats including mangroves and coastal salt marshes (salt flats) are also considered as separate assets within this Assessment. Figure 3-5 White-bellied sea-eagle (Haliaeetus leucogaster) in a wetland in northern Australia Photo: CSIRO The Roper catchment has two nationally significant wetlands listed under the DIWA: Limmen Bight (Port Roper) Tidal Wetlands System and Mataranka Thermal Pools (Figure 3-6) (Department of Agriculture‚ Water and the Environment, 2021b). There are no Ramsar-listed wetlands within the Roper catchment. The Limmen Bight (Port Roper) Tidal Wetlands System is approximately 185,000 ha and is located at the mouth of the Roper River (Department of Agriculture‚ Water and the Environment, 2021b). This wetland system includes intertidal mud flats, saline coastal flats and estuaries, and has a high volume of freshwater inflows (McJannet et al., 2009). The area includes the Limmen National Park south of the Roper River, and the South East Arnhem Land Indigenous Protected Area to the north of the Roper River. The area is considered a site of conservation significance, and is important for seabirds, waterbirds and migratory shorebirds (Smyth and Turner, 2019). It also supports commercial fisheries for prawns, mud crabs and barramundi (Smyth and Turner, 2019). Traditional practices are still carried out in the South East Arnhem Land Indigenous Protected Area (Gambold, 2015). The Mataranka Thermal Pools in Elsey National Park in the upper reaches of the Roper catchment are a series of permanent, groundwater-connected thermal springs, fed via flows through the Tindall Limestone Aquifer (Figure 3-6). The thermal pools are inchannel habitat of less than 10 ha in total area (McJannet et al., 2009). For the purpose of this Assessment, the Mataranka Thermal Pools are covered within the waterholes and GDE assets. As well as these two nationally significant wetlands, there are several floodplain areas within the Roper catchment (see ‘land subject to inundation’, Figure 3-6). These floodplain areas flood during For more information on this figure please contact CSIRO on enquiries@csiro.au the wet season, replenishing associated semi-permanent and permanent wetlands. Significant numbers of freshwater floodplains occur in association with the rivers and creeks within the Roper catchment, particularly the Roper, Hodgson, Jalboi and Wilton rivers, and on the Flying Fox, Maiwok, Birdum, Jasper, Horse and Showell creeks (Figure 3-6). Figure 3-6 Land subject to inundation (potential floodplain wetlands) and nationally important wetlands (DIWA) in the Roper catchment DIWA = Directory of Important Wetlands in Australia Dataset: Geoscience Australia (2017); Department of the Environment and Energy (2010) For more information on this figure please contact CSIRO on enquiries@csiro.au Barramundi Barramundi are arguably the most important fish species to commercial, recreational and Indigenous subsistence fisheries throughout Australia’s wet-dry tropics. Barramundi make up a substantial component of the total commercial fish catch in northern Australia (Savage and Hobsbawn, 2015). In 2013–14, barramundi comprised 28% of the $31 million wild-caught fishery production in the NT. Commercial catch-per-unit-effort in the NT has increased from about 7 kg per 100 m of net per day in the early 1980s to over 30 kg per 100 m of net per day in the 2010s (Northern Territory Government, 2018). The commercial and recreational catches make up the largest proportions of all catches in the NT, though the Indigenous artisanal catch is significant in some years. Barramundi are also a fish of cultural significance for the Indigenous community as well as being an important food source (Jackson et al., 2012). The movements of barramundi between habitats are indicators of the change in season to Indigenous communities across tropical Australia (Green et al., 2010). Their movements are related to habitat requirements during their life cycle and the reliance of barramundi on seasonal variation in river flows to access these habitats. In the NT, the Indigenous catch of barramundi in the study area is less certain than other fisheries. Barramundi life history renders them critically dependent on river flows (Tanimoto et al., 2012). Large females (older fish) and smaller males (younger fish) reside in estuarine and littoral coastal habitats. Mating and spawning occur in the lower estuary during the later dry season to early wet season and new recruits move into supra-littoral and freshwater habitats; with coastal salt flat, floodplain and palustrine habitats dependent on overbank flows for maintenance and connectivity (Crook et al., 2016; Russell and Garrett, 1983, 1985). Barramundi occupy relatively pristine habitats in both freshwater and estuarine reaches of the Roper River and are abundant in the river. The remote location of the Roper catchment is linked to low numbers of reports of barramundi in freshwater reaches of the river and its tributaries (e.g. from recreational fishers). However, given the commercial catch and the known ecology of the species from other catchments it is likely that the Roper catchment represents an important system for barramundi (Crook et al., 2016; Dostine and Crook, 2016). Grunters (Family: Terapontidae) Grunters inhabit riverine, estuarine and marine waters and in northern Australia there are a total of 37 species of grunter from 11 genera, with the most species-rich genera being Hephaestus, Scortum, Syncomistes and Terapon (Figure 3-7). Many grunter species spend their entire lives in fresh water, while other species inhabit marine or estuarine waters, only sometimes venturing into fresh water (Pusey et al., 2004). The Terapontidae are a perciform (‘perch like’) family of fishes of medium diversity, restricted to the Indo-Pacific region. They are characterised by a single long-based dorsal fin, which has a notch marking the boundary between the spiny and soft-rayed portions, and are a soniferous family (i.e. they can both vocalise and hear well), and thus may be sensitive to noise (Smott et al., 2018). One of the most ubiquitous species is the sooty grunters (Hephaestus fuliginosus). Sooty grunters are omnivorous and their diet is diverse in composition, containing terrestrial insects and vegetation, fish, aquatic insect larvae, macrocrustacea (shrimps and prawns) and aquatic vegetation. Sooty grunters switch diet from being insectivorous while juvenile to being top-level predators as adults, often feeding on smaller fish as well as juvenile grunters. Juvenile grunters are often associated with flowing water, suggesting that water harvesting that reduces or ceases flow could pose a threat. Tree root masses and undercut banks are also important microhabitat, especially for adult fish (Pusey et al., 2004). Grunters prefer medium to high oxygen levels as well as medium to low salinity (Hogan and Nicholson, 1987). Grunters will move out of the dry-season refugial habitats and into ephemeral wet-season habitats for spawning (Bishop et al., 1990), with juveniles known to swim up to 7 km. The sooty grunter are an important recreational species, with environmental flow being managed to maintain suitable habitat conditions (Chan et al., 2012). Because grunters are omnivorous and able to integrate many sources of food, as well as having a high overall biomass, they are an important link in the overall food chain. They bridge lower trophic levels with top-level predators, such as long tom (Strongylura krefftii) or crocodiles. Grunters are also important species for Indigenous people in northern Australia, both culturally (Finn and Jackson, 2011; Jackson et al., 2011) and as a food source (Naughton et al., 1986). There are seven species of grunters in the Roper catchment (Figure 3-7): spangled grunter (Leiopotherapon unicolor), barred grunter (Amniataba percoides), sooty grunter, Gulf grunter (Scortum ogilbyi) and estuarine trumpeter (Pelates quadrilineatus). Of these, sooty grunters are the key species for recreational and cultural purposes (Chan et al., 2012). In the Roper catchment, grunters are likely widespread with headwaters being spawning and nursery grounds, as well as habitat for adults of the smaller species (spangled grunter for example). Waterholes on the main stem of the Roper River represent habitat for adult grunters. Figure 3-7 Grunters in the Roper catchment Data source: ALA (2021) Freshwater turtles Freshwater turtles are one of the world’s more endangered taxonomic groups, with 52% of global species extinct or threatened (Böhm M et al., 2013; Van Dijk et al., 2014). Freshwater turtles in Australia can be divided into three families: Chelidae (32 species), Trionychidae (two species), and Carettochelyidae (one species) (Georges and Thomson, 2010). Chelids, members of the Chelidae family, are highly aquatic species. They have webbed feet and can stayed submerged in water for long periods of time. Chelids retract their necks sideways into their shells and their dietary habits For more information on this figure please contact CSIRO on enquiries@csiro.au vary between genera. Long-necked species, such as Chelodina spp., are largely carnivorous, feeding on fish, invertebrates and gastropods (Legler, 1982; Thomson, 2000); while short-necked species, such as Elseya spp., are herbivorous or specialised to eat fruits (Kennett and Russell- Smith, 1993). Freshwater turtles depend upon flooded wetland systems for breeding, nesting, food provision and refuge. Changes to regional hydrology, habitat loss and climate change are some of their key threatening processes (Stanford et al., 2020). In northern Australia, turtles occupy a range of aquatic habitats, including both river and floodplain wetland habitats such as main channels, waterholes, floodplain wetlands and oxbow lakes (Cann and Sadlier, 2017; Thomson, 2000). Many of the turtle species in northern Australia have developed adaptive traits to survive the inter-annual variation between the wet and dry seasons, such as the emergence of hatching with the wet-season onset (Cann and Sadlier, 2017). During the dry season, the movements of freshwater turtles on and off the floodplain are limited, making them more vulnerable to changes in water quality, invasive species and habitat degradation (Cann and Sadlier, 2017; Doupe et al., 2009). Australian freshwater turtles are of both ecological and cultural significance in Australia. This includes the consumption of some species by Indigenous peoples as a seasonal source of protein (Jackson et al., 2012). A recent collaboration between Yangbala Rangers and the Atlas of Living Australia provided shared cultural values, and threats, ecological knowledge and distribution of freshwater turtles in the Roper catchment. This is the first time that regionally specific Indigenous observational occurrence data and Indigenous historical knowledge are included in the Atlas of Living Australia (Daniels et al., 2022). Indigenous people have widespread connections to freshwater turtles through songlines and ceremonies and certain people have roles as custodians and caretakers according to the kinship system. Knowledge holders described seasonal knowledge and indicators that related to freshwater turtle hunting, behaviour, diet and physiology, including aestivation, fatness and breeding cycles. For example, knowledge holders said the dry (cold) season is the time to hunt for northern snake-necked turtle (Chelodina oblonga oblonga; previously known as Chelodina rugosa). The main threats to the freshwater turtles, identified by the Indigenous peoples, were natural predators (including birds of prey (such as eagles and hawks), crocodiles, goannas and dingoes), feral animals (such as pigs, buffalo, horses, donkeys, cattle and cane toads) and climate change (e.g. lower rainfall) (Russell et al., 2021). There are ten species of freshwater turtles described in the NT (Northern Territory Government, 2017). In the Roper catchment there are records of five freshwater turtle species: Gulf snapping turtle, northern snapping turtle (Elseya dentata), northern snake-necked turtle, Cann’s snake- necked turtle (Chelodina canni) and red-bellied short-necked turtles (Emydura subglobossa) (Figure 3-8). Until the recent collaboration with Yangbala Rangers, records for this area were sparse compared to many other regions of Australia. These turtles occur in different habitats, from permanently flowing riverine habitats to lakes, billabongs and swamps, from the Roper River mouth to Mataranka, but more surveys are required to assess their current distribution and conservation status in the study area. Currently all five species are listed as Least concern by the Northern Territory Government; however, the northern snapping turtle is listed federally as Endangered by the EPBC Act. Figure 3-8 Distribution of freshwater turtles within the Roper catchment The freshwater turtle dataset was created from a collaboration between Ngukurr Yangbala Rangers, members of Ngukurr and Numbulwar communities (South East Arnhem Land), Macquarie University ecologists and the Atlas of Living Australia through a series of mapping workshops and interviews to record local knowledge of the distribution in the South East Arnhem Land Indigenous Protected Area of the freshwater turtles. Elseya lavarackorum distribution modelled through Species of National Environmental Significance (SNES; Department of Agriculture, Water and the Environment (2019c)). Data sources: ALA (2022); Department of Agriculture, Water and the Environment (2019c); Department of Environment, Parks and Water Security (2019a) For more information on this figure please contact CSIRO on enquiries@csiro.au. Due to the richness of natural resources across parts of the Roper catchment, this area was probably seasonally exploited in a fisher, hunter, gatherer economy, allowing large groups of people to gather for ceremony and other purposes. This is represented in rock art sites in the region, with at least one known site showcasing turtles (David et al., 2017; Earth Sea Heritage Surveys, 2013). Note that the recognised Australian distribution of the pig-nosed turtle (Carettochelys insculpta) occurs in the western and northern draining catchments of the Gulf of Carpentaria in the NT and the species has been reported in the Roper catchment, although not confirmed (Georges et al., 2008). Similarly, the sandstone snake-necked turtle (Chelodina burrungandjii), currently listed as data deficient by the Northern Territory Government, is known to occur in the Wilton River, a tributary of the Roper River (Thomson et al., 2000). Waterbirds: colonial and semi-colonial nesting wading The colonial and semi-colonial nesting, wading waterbirds (‘colonial waders’) group comprises wading waterbird species that have a high level of dependence on water for breeding, including requirements for flood timing, extent, duration, depth, vegetation type and vegetation condition. In northern Australia, this group comprises 21 species from five families, including ibis, spoonbills, herons, egrets, avocets, stilts, storks and cranes. The species in this group are often easily detectable when breeding and relatively good datasets are available for most, compared to other species or groups. The species in this group are often dependent on specific important breeding sites (Arthur et al., 2012). Ibis, spoonbills, herons, egrets, avocets and stilts nest in loose groups or dense colonies of hundreds of birds to tens of thousands of birds in specific vegetation types and locations, over or adjacent to water (Bino et al., 2014). Storks (such as the black-necked stork; Ephippiorhynchus asiaticus) and cranes including the brolga (Antigone rubicunda) and sarus crane (Antigone antigone) usually nest independently, but loose, widely spaced groups of nests may occur in suitable habitat. Species in this group may travel significant distances to use these sites, ranging up to thousands of kilometres (McGinness et al., 2019), and nesting events can last several months, depending on inundation conditions (Kingsford et al., 2012). Species in this group usually have a mixed diet including fish, frogs, crustaceans and insects, and use foraging methods such as walking, stalking and striking to catch their prey. Colonial and semi-colonial waders generally prefer shallow water or damp sediment with medium to low-density vegetation for foraging (Garnett et al., 2015). These species are typically nomadic or partially migratory but may spend long periods in particular locations when conditions are suitable. From the colonial and semi-colonial nesting waders group, the royal spoonbill (Platalea regia) (Figure 3-9) is a large wading species highly adapted to foraging in shallow wetlands (Marchant and Higgins, 1990). This species requires water and water-dependent vegetation for feeding, nesting, refuge, roosting and movement habitat (e.g. stopover habitat for longer distance trips) (Marchant and Higgins, 1990). Spoonbills nest in loose colonies, usually in vegetation surrounded by water, including reedbeds, semi-aquatic shrubs and trees. They often nest adjacent to colonies of other species in the group. Colonial and semi-colonial nesting waders, including the royal spoonbill, are found widely throughout the Roper catchment. The large wetlands and extensive mangroves, including the areas throughout Limmen National Park, support a range of colonial and semi-colonial nesting waders (Delaney, 2012). Aerial surveys by Chatto (2006) found large numbers of egrets (including Egretta spp.) and the red-necked avocet (Recurvirostra novaehollandiae). The mangrove habitats along the north side of the Roper River support significant colonies of great egret (Ardea alba), intermediate egret (Ardea intermedia), little egret (E. garzetta) and pied heron (E. picata), with the nankeen night-heron (Nycticorax caledonicus) found along the banks of the river in the mangrove habitat as far as Ngukurr (Smyth and Turner, 2019). The Roper River is also considered to be a major breeding area for brolgas during the wet season (Chatto, 2006). Permanent waterholes and wetlands around Mataranka also support a variety of colonial and semi-colonial nesting waders. Figure 3-9 Royal spoonbills are a representative species of the colonial and semi-colonial nesting waders functional group Photo shows individuals at the nest. Photo: CSIRO For more information on this figure please contact CSIRO on enquiries@csiro.au 3.2.3 Ecological assets from marine systems The marine and estuarine habitats of northern Australia include some of the most important, extensive and intact habitats of their type in Australia, many of which are of national significance. Marine habitats in northern Australia are vital for supporting important fisheries including the common banana prawn, mud crab and barramundi, as well as for biodiversity more generally, including waterbirds and marine mammals and turtles. In addition, the natural waterways of the sparsely populated catchments support globally significant stronghold populations of endangered and endemic species (e.g. sharks and rays) that use both marine and freshwater habitats. This section provides a synthesis of the prioritised assets relevant to marine sections of the Assessment catchments. Table 3-1 presents the full list of assets used in the Roper catchment ecology assessment and this section provides information on a sample of these as relevant to the marine systems of this catchment. The Roper catchment marine region as considered here is an area south of Groote Eylandt that depends upon recruitment from littoral habitats from the Roper to the mainland coast to the west of Groote Eylandt. Banana prawns Banana prawns are large-bodied decapod crustaceans around 80 g in size of the family Penaeidae that are found throughout the Indo-West Pacific. They are a prized fishery target species throughout their geographic distribution. Two species of banana prawns are found in Australia, the common banana prawn and the redleg banana prawn. Both banana prawn species are globally widespread throughout the Indian Ocean and south-east Asian and west Pacific coastal habitats. In Australia, common banana prawns inhabit tropical and subtropical coastal waters (Grey et al., 1983). In contrast, the Joseph Bonaparte Gulf and western Tiwi Island region in north-west Australia are the south-eastern limit of the worldwide distribution of redleg banana prawns (Grey et al., 1983). Common banana prawns are prolific in the western Gulf of Carpentaria, with significant commercial catches taken adjacent to their inshore estuarine habitats (Staples et al., 1985). Banana prawns support an approximate 4942 t ‘sub-fishery’ component (recent 10-year mean) of the Northern Prawn Fishery (NPF) (worth about $70–80 million annually) (Laird, 2021). The major portion of the common banana prawn catch is taken in the eastern Gulf of Carpentaria; however, significant catches are taken offshore from the Roper River (Laird, 2021). The influence of rainfall and runoff from western Gulf of Carpentaria catchments on banana prawn catches is less clear than for eastern catchments and requires further investigation, though seasonal rainfall and prevailing winds are positively correlated with catch (Vance et al., 1985, 2003). Using commercial catch as a measure of population abundance, large flood flows cue the prolific population of juvenile banana prawns to emigrate en masse to the near-shore and offshore zones where they rely on marine habitats for enhanced growth and survival (Broadley et al., 2020; Duggan et al., 2019; Lucas et al., 1979). Adult banana prawn distribution is adjacent to their juvenile estuarine mangrove habitats (Staples et al., 1985; Zhou et al., 2015). Adult common banana prawns occupy soft-sediment substrates in relatively shallow waters within the south-west, south-east and eastern Gulf of Carpentaria, and along the Top End/Arnhem Land coastline. Banana prawns are managed by limited effort (licence to fish) and by spatial and temporal closures. The fishing season opens on 1 April annually and continues until catch rates decline to a trigger level defined in the Northern Prawn Fishery Harvest Strategy (AFMA, 2022). Adult common banana prawns live and spawn offshore in waters 10–30 m deep, the larvae and postlarvae move by drift inshore to settle in the mangrove forest and mudbank matrix in estuarine mangrove habitats (Crocos and Kerr, 1983; Staples, 1980; Vance et al., 1998). Each of the major rivers along the south-west Gulf of Carpentaria coastline from Blue Mud Bay to the south-east Gulf of Carpentaria support abundant populations of juvenile banana prawns (Staples, 1979). Common banana prawns are found at highest densities offshore from the Roper River in relatively shallow waters, as well as south of Groote Eylandt in deeper water (Figure 3-10). Common banana prawns are abundant elsewhere in the western Gulf of Carpentaria in Blue Mud Bay north of Groote Eylandt (north of the map extent in Figure 3-10) and offshore of the McArthur River to the south-east of the Roper catchment marine region. The Roper catchment marine region lies within the southern portion of the ‘Groote’ NPF statistical region of the Gulf of Carpentaria (adjacent to the Blue Mud Bay and Roper River coasts). This statistical region accounts for about 2% (about 95 t – 16-year mean catch) of the total NPF banana prawn catch (Laird, 2021). In all locations, the highest abundances of banana prawns were caught inshore in about 15–20 m depth, in proximity to the river estuaries (Zhou et al., 2015). In the 1970s, the use of mangrove habitats by juvenile banana prawns within the Roper River estuary was documented by Staples (1979) using a float plane to access a series of rivers in the region. However, the remoteness of the river systems in the western Gulf of Carpentaria render both the estuarine habitats and their fish and crustacean fauna poorly studied. Knowledge of estuarine banana prawn habitats from other Gulf of Carpentaria rivers showed that the mangrove forest and creek mudbank habitats (indicated as juvenile habitat in Figure 3-10) are critical for juvenile banana prawn survival and growth (Staples, 1979; Vance et al., 1990). These habitats are prolific within the estuaries of many rivers in the western Gulf of Carpentaria, including the Roper River, as well as other rivers along the south-west Gulf of Carpentaria coastline, such as the Limmen Bight and McArthur River (Duke et al., 2017). Figure 3-10 Fisheries catch of banana prawns and their habitat in the Roper catchment marine region Banana prawn juveniles use the estuary and adult prawns are caught offshore in water about 10–20 m deep in the marine habitat. Units are kilograms as total catches for the 10-year period 2011 to 2020. Data sources: Kenyon et al. (2022); Staples (1979) For more information on this figure please contact CSIRO on enquiries@csiro.au Mud crabs Mud crabs are a large-bodied, large-clawed, short-lived, fast-growing decapod crustacean (>200 mm carapace width) that inhabit the estuarine and shallow subtidal community along tropical and subtropical coastlines, especially mangrove-dominated habitats. They are targeted throughout their range as a commercial, recreational and Indigenous fishery resource and a prized table species (commercial catch 40,000 t worldwide in 2012) (Alberts-Hubatsch et al., 2016). Two species of mud crab are found in tropical Australia, Scylla serrata and S. olivacea (Alberts-Hubatsch et al., 2016; Robins et al., 2020). Mud crabs are distributed across the Indo-Pacific region; though in Australia, S. serrata is the dominant commercial species by abundance (Robins et al., 2020). Scylla olivacea is found only in the north-east Gulf of Carpentaria in the Weipa region (Alberts- Hubatsch et al., 2016; Robins et al., 2020). The combined NT and Queensland Gulf of Carpentaria mud crab catch contributed about 25% of the reported mud crab commercial harvest in Australia between 2008 and 2017. The NT crab catch in 2018–19 was 270 t valued at $7,881,000, while the Queensland crab catch was 1949 t valued at $19,825,000 (all crab species, Steven et al., 2021). At the Sydney Fish Market, the price for mud crabs averaged about $34/kg in 2018–19, making them a high-value regional resource (Robins et al., 2020). The mud crab’s high fecundity, high natural mortality and relatively short life span suggest that they are a moderately resilient species suitable for sustainable harvest. The high market price commanded by mud crabs supports their fishery within, and transport from, remote coastal locations in tropical Australia, including the Gulf of Carpentaria region. Mud crabs occupy mangrove forest and nearby shallow subtidal habitats within estuarine and coastal ecosystems (Alberts-Hubatsch et al., 2016) (example habitat shown in Figure 3-11), hence they use the estuaries and shallow-water coasts in the Gulf of Carpentaria as habitat. Mud crabs are an important ecological species, being both predator and prey in the coastal ecosystem. As small juveniles, mud crabs are detritivores, as large juveniles and as adults they are benthic predators feeding on crustaceans, molluscs and fish. Estimates suggest that the mud crab population consumes 650 kg biomass per ha per year in the mangrove forest and 2100 kg biomass per ha per year in mangrove fringe habitat (Alberts-Hubatsch et al., 2016). Mud crabs dig burrows to rest during the day, reworking mud substrates within mangrove forests and mudbanks. They play a significant trophic role in mangrove ecosystems. Figure 3-11 Mangrove and intertidal habitat associated with mud crabs in northern Australia Photo: CSIRO Mud crabs demonstrate a larval life-history strategy (see Robins et al. (2020) for a recent comprehensive review): females migrate offshore to spawn after the adult crabs mate in the estuary (September to November, larvae require marine salinity) (Hill, 1975, 1994; Meynecke et al., 2010; Welch et al., 2014). Their larvae transform to megalopae (the final larval stage) that move by drift inshore where they settle as benthic juveniles in estuarine mangrove and mudflat habitats (Alberts-Hubatsch et al., 2016; Meynecke et al., 2010; Robins et al., 2020). The larval form facilitates not only migration as crabs grow to the juvenile stage (ontogenetic migration) and settle to their inshore habitats, but long-distance dispersal and genetic mixing (Gopurenko and Hughes, 2002; Gopurenko et al., 2003; Robins et al., 2020). Initial recruitment to inshore habitats occurs at the mangrove forest fringe, while as crabs grow, their dependence on estuarine mangroves declines (Alberts-Hubatsch et al., 2014). Mud crabs remain in the estuary for several years as sub- adults and adults, before the females alone emigrate to spawn (Hill, 1994). Regionally, the annual wet season and subsequent runoff is a significant determinant of their recruitment strength and total catch (possibly lagged by 1 to 2 years) in the estuary and near-shore zone (Meynecke and Lee, 2011; Meynecke et al., 2010). However, recent analyses of Gulf of Carpentaria catches support the notion of river flow enhancing catch, but also show high air temperature over the wet season as a dominant negative influence on mud crab abundance within the Roper River and southern Gulf of Carpentaria estuarine habitats (Robins et al., 2020). For more information on this figure please contact CSIRO on enquiries@csiro.au From 2006 to 2018, the average harvest of mud crabs for the Roper catchment marine region was 71 t (an average 35% of the harvest from the Northern Territory Western Gulf of Carpentaria Mud Crab Fishery) (Robins et al., 2020). The Roper catchment marine region had a high variation in catch: a minimum catch of 3.3 t in 2016 and a maximum catch of 123.5 t in 2009 (Robins et al., 2020). Robins et al. (2020) conducted a recent comprehensive analysis of the effect of environmental drivers on Gulf of Carpentaria mud crab catches and found that within the western and south-western regions of the Gulf of Carpentaria, heat, evaporation, precipitation, water stress and sea level, as well as hemisphere-wide phenomena create a high-stress environment for mud crabs (i.e. within the Roper and McArthur river estuaries). 3.2.4 Ecological assets from terrestrial systems The terrestrial habitats of northern Australia include a range of varied and significant habitat types. While much of the tropics of northern Australia is savanna, eucalypt forest and grasslands, other habitats include riparian and floodplain communities, and GDEs including aquatic, terrestrial and subterranean habitats. Many of these are highly dependent upon fresh water from rivers and can also be supported by groundwater discharge for their persistence and condition. Surface water dependent vegetation Across much of northern Australia, terrestrial vegetation survives on water derived from local rainfall that recharges soils during the wet season and can be accessed by the root systems within unsaturated soils throughout the year. Terrestrial vegetation that receives extra water (i.e. in addition to local rainfall, for example, through recharge from flood waters or by accessing shallow groundwater), often provides a lush green and productive forest ecosystem (high diversity, dense tree cover) within an otherwise drier or more sparsely vegetated savanna environment (e.g. Pettit et al., 2016). This is referred to as surface water dependent vegetation. While water availability influences the distribution of savanna versus forest ecosystems across the northern Australia landscape, their distribution is also linked to fire regime, nutrient availability, soil type and herbivory (Murphy and Bowman, 2012). Terrestrial vegetation that receives extra water may contain unique species (e.g. Critically endangered Carpentarian rock-rat (Zyzomys palatalis), unique to monsoon forests, Crowley (2010)) and provide critical habitat for fauna (e.g. Melaleuca forests in the NT support many nationally significant rookeries for waterbirds (Woinarski, 2004)). Such habitats often occur along rivers and floodplains, fringing wetlands and springs or where the depth to groundwater is within reach of the roots. Vegetation naturally inhabits and thrives in niches in the environment that provide the right combination of water conditions including: • surface water depth (during low and high flows) • groundwater depth • timing and flood frequency (return interval) • flood duration. The optimal water regime will vary for different climate conditions (rainfall regime), site conditions (soil type and water availability) and vegetation types. The water regime supports vegetation survival, growth, flowering and fruiting, germination and successful establishment of new saplings for the diversity of ecosystem species and maintains their functions and services. Vegetation is unlikely to be able to adapt to changes in water availability outside natural variation. It has some inbuilt resilience to natural changes in water availability, but prolonged change is likely to result in dieback after some lag period and shift in ecosystem structure and function (e.g. Mitchell et al., 2016). Terrestrial vegetation that requires surface water inundation and/or access to groundwater is at risk from water resource development if the natural surface water and groundwater regimes are modified beyond some limit. In northern Australia, these ecosystems provide food and habitat for high levels of biodiversity (e.g. for migratory waterbirds, flying-foxes, crocodiles and honeyeaters), play a role in nutrient cycling and provide buffering against erosion. Three vegetation communities are considered as part of the ecology assessment as they are communities that are regularly inundated with surface water during wet seasons, namely paperbark swamps, river red gum and monsoon vine forest (Figure 3-12). These are described below with further information in the companion technical report on ecological assets (Stratford et al., 2022). Paperbark is a term commonly used to describe a range of Melaleuca species that have a distinctive papery bark texture. Some paperbark species occur in low-lying areas that are seasonally inundated with fresh water (Department of Environment and Science Queensland, 2013). Many paperbark species co-occur with eucalypt species in riparian and floodplain tree swamps (Department of Environment and Science Queensland, 2013). For the purpose of this assessment, a ‘paperbark swamp’ refers to the non-tidal coastal and subcoastal swamp that are dominated by Melaleuca species with papery-textured bark (Department of Environment and Science Queensland, 2013). River red gum commonly line permanent or seasonal rivers and sometimes form forests over floodplains (Costermans, 1981) that are subject to frequent or periodic flooding. Flooding requirements for maintaining healthy river red gum have been estimated for various floodplain forests and riparian woodlands in the Murray–Darling Basin (MDB) ranging from every 1 to 3 years for 2 to 8 months (Rogers and Ralph, 2010) to every 3 to 5 years for up to 2 months (Wen et al., 2009). River red gum may require flood to induce seed fall (George, 2004), but excessive flooding can destroy seeds (Rogers and Ralph, 2010). Note that these flooding relationships exist for trees found in the MDB where there has been extensive research completed on maintaining this ecosystem type; however, they cannot be directly extrapolated to the different hydrological-soil- climate conditions of northern Australia. Specific water requirements for river red gum and subspecies found in northern Australia are unknown. Monsoon vine forest can be found in tropical and subtropical regions of northern Australia, with patches spanning the NT, Queensland and WA. While generally falling under the umbrella term ‘rainforest’ with its closed canopy and high leaf cover exceeding 70% (Stork et al., 2008), it can be further characterised by canopy height, leaf size, proximity to permanent moist soils and species composition. This forest type is typically found in areas of 600–2000 mm mean annual rainfall (Bowman, 2000). Most monsoon vine forests seem limited to areas with permanent soil water, such as creek lines, springs and seeps, and are thought to be remnants of a wetter period during Australia’s geological history; changes in climate, fire regime and water availability has restricted their distribution to small pockets across northern Australia of less than several hectares (Bowman, 2000). However, the hydrological and geomorphic environments of these ecosystem communities are poorly understood, and while monsoon vine forests can typically be found in areas that offer fire protection, such as boulder outcrops and areas of high soil water, a change in water availability may make them more prone to fire (Larsen et al., 2016; Russell‐Smith, 1991). Figure 3-12 Locations of observed selected surface water dependent vegetation types in the Roper catchment Species within each vegetation type are provided in the companion technical report on ecological assets (Stratford et al., 2022). GDE = groundwater-dependent ecosystem, MVF = monsoon vine forest Dataset: ALA (2021); Department of Environment, Parks and Water Security (2000) For more information on this figure please contact CSIRO on enquiries@csiro.au 3.2.5 Environmental protection There are a number of both aquatic and terrestrial species in the Roper catchment currently listed as Critically endangered, Endangered and Vulnerable under the EPBC Act and by the Northern Territory Government’s wildlife classification system, which is based on the International Union for Conservation of Nature (IUCN) Red List categories and criteria (Figure 3-13). Figure 3-13 Distribution of species listed under the EPBC Act (Cth) and by the Northern Territory Government in the Roper catchment For more information on this figure please contact CSIRO on enquiries@csiro.au If a proposed development is predicted to have a significant impact on a matter of national environmental significance (e.g. populations of a nationally listed species, community, migratory species or wetlands of importance) it would require approval to proceed under the EPBC Act (Table 3-2). This approval is required irrespective of local government policies. The Commonwealth’s Protected Matters Search Tool lists 43 Threatened species for the Roper catchment, 4 of which are listed as Critically endangered. Also listed are 47 migratory species. Table 3-2 Definition of threatened categories under the EPBC Act (Cth) and the Northern Territory wildlife classification system For more information on this figure, table or equation please contact CSIRO on enquiries@csiro.au †The NT wildlife classification categories are based on the IUCN Red List categories and criteria. An extract of each category is presented here. For the full definition see https://nt.gov.au/__data/assets/pdf_file/0010/192538/red-list-guidelines.pdf. 3.3 Demographic and economic profile 3.3.1 Introduction This chapter describes the current social and economic characteristics of the Roper catchment in terms of the demographics of local communities (Section 3.3.2), the current industries and land use (Section 3.3.3), and the existing infrastructure of transport networks, supply chains, utilities and community infrastructure (Section 3.3.4). Together these characteristics describe the built and human resources that would serve as the foundation upon which any new development in the Roper catchment would be built. Unless otherwise stated, the material in this section is based on findings described in the companion technical report on agricultural viability and social economics (Stokes et al., 2023). 3.3.2 Demographics The Roper catchment comprises around half of the Roper Gulf Regional Council local government area together with small parts of a number of other adjacent local government areas, including Katherine Town Council, West Arnhem Regional Council, East Arnhem Regional Council and Victoria Daly Regional Council. At the state/territory level the catchment includes the majority of the electoral division of Arnhem and a small part of a number of other electoral divisions, including Katherine, Arafura, Mulka and Gwoja. At the federal level the catchment forms a part of the Division of Lingiari (which encompasses the majority of the NT, excluding the Division of Solomon that covers an area near Darwin). The population density of the Roper catchment is extremely low at one person per 32.6 km2, which is about five times lower than the NT, and 100 times lower than Australia as a whole. The region contains no large urban areas (population >10,000 people), however, there are a number of small towns and communities within the catchment including Barunga, Beswick, Bulman, Daly Waters, Larrimah, Mataranka (the regional centre), Minyerri and Ngukurr. The only one of these settlements with a population greater than 1000 is Ngukurr (population 1149 as at the 2016 Census). Katherine (population 6303 in 2016) is the closest urban service centre and is located about 100 km north-west of Mataranka, just outside the catchment. The nearest major city and population centre is the NT capital of Darwin (population of Greater Darwin area was 136,828 in 2016), approximately 420 km from Mataranka. The demographic profile of the catchment, based on data from the 2016, 2011 and 2006 censuses is shown in Table 3-3. The Australian Bureau of Statistics (ABS) reports statistics by defined statistical geographic regions (such as the nested hierarchy of statistical areas), but none of those regions closely approximate the Roper catchment. Instead, data are shown for: (i) Elsey (ABS Statistical Area Level 2 (SA2) region 702051065), being the single region which most closely approximates the catchment boundary (Figure 3-14); and (ii) estimated data based on combining the appropriate portions of a number of ABS regions to best match the actual spatial coverage of the catchment (62.2% of Elsey SA2 region, 19.0% of Gulf SA2 region, plus small proportions (each less than 2%) of the SA2 regions of East Arnhem, Katherine, Victoria River and West Arnhem). The typical resident of the region is younger, poorer and more likely to identify as Indigenous than the typical resident of the NT and of Australia as a whole. The population is predominantly younger (median age less than 30) than is typical compared to the NT and to the country as a whole (median age more than 30), however, the trend from 2011 to 2016 suggests that the median age is moving towards the NT and national averages. The population contains a much larger proportion of Indigenous people (more than 70%), compared to the NT (25.5%) and the country overall (less than 3%), and the median household income was considerably below the average for the NT and for the country as a whole in 2016. Furthermore, the proportion of households on low incomes (less than $650/week) was far higher, and the proportion on high incomes (more than $3000/week) far lower than the proportion for the NT and for the country as a whole. Figure 3-14 Boundaries of the Australian Bureau of Statistics Statistical Area Level 4 (SA4) and Statistical Area Level 2 (SA2) regions used for demographic data in this Assessment Table 3-3 Major demographic indicators for the Roper catchment For more information on this figure or table please contact CSIRO on enquiries@csiro.au Se-R-505_Map_Australia_Roper_tourism_SA2_v3 For more information on this figure or equation please contact CSIRO on enquiries@csiro.au For more information on this figure or table please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the catchment boundary. Source: ABS (2006, 2011, 2016) census data The Roper catchment falls within the 1st decile for each of the Socio-economic Indexes for Areas (SEIFA) metrics (Table 3-4), indicating the region is scoring below 90% of the rest of the country on each of the measures. When considering the various SA2 regions that fall within the catchment boundary, virtually all (West Arnhem, Elsey, Gulf, Victoria River) individually rank within the 1st decile for each of the four measures. Only the Katherine region (less than 2% of which falls within the Roper catchment border) avoids this lowest decile for all measures (ranging from 3rd to 6th decile) while the East Arnhem region (less than 1% of which falls within the Roper catchment border) ranks in the 2nd decile for the Index of Education and Occupation (IEO). Table 3-4 SEIFA scores of relative socio-economic advantage for the Roper catchment Scores are relativised to a national mean of 1000, with higher scores indicating greater advantage. For more information on this figure or table please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the catchment boundary. ‡Accessibility and Remoteness Index of Australia Score. 1Based on both the incidence of advantage and disadvantage. 2Based purely on indicators of disadvantage. Source: ABS (2016) 3.3.3 Current industries and land use Employment The economic structure of the Roper catchment differs substantially from that of the NT and Australia as a whole. The proportion of the adult population (aged 15 and older) within the labour force is far smaller (see participation rates within Table 3-5), indicating that a large proportion of the potential workforce is unable or unwilling to seek work. Furthermore, the unemployment rates are far higher than the NT and national averages (see unemployment rates within Table 3-5), indicating that of those who are willing and able to seek work a larger proportion have been unable to find work. There are also noticeable differences in the industries providing the most jobs within the region (Table 3-5). ‘Education and training’ and ‘Health care and social assistance’ are important employers in the region and nationally, but while ‘Retail trade’, ‘Construction’ and ‘Professional, scientific and technical services’ feature within the top five industries by employment across the nation on average, they are far less significant within the Roper catchment. Similar to the NT as a whole, ‘Public administration and safety’ and ‘Other services’ are relatively more important to the employment prospects of workers within this region compared to the national average. However (and of particular relevance to this Assessment), ‘Agriculture, forestry and fishing’ features strongly within the top five industries for the Roper catchment, and furthermore, the importance of the sector has been growing over time when results of the previous censuses are considered. Over the last three censuses (2006, 2011 and 2016) the percentage of employment from the agricultural sector nationally has been reported as 3.1%, 2.5% and 2.5%, respectively, and for the NT, 2.4%, 1.9% and 2.0%, respectively, over the same years. That is, the industry proportion of employment in the sector has been small and fairly flat. In contrast, the importance of agricultural employment within the Roper catchment is large and growing, having provided 12.2% of employment in 2006, 13.5% in 2011 and 14.0% in 2016. The structural differences in this region compared to elsewhere can have a significant impact on the regional economic benefits that can result from development projects initiated within the region compared to development projects that may be initiated elsewhere. Table 3-5 Key employment data for the Roper catchment For more information on this figure or table please contact CSIRO on enquiries@csiro.au For more information on this figure or table please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the catchment boundary. Source: ABS (2006, 2011, 2016) census data Land use The Roper catchment covers an area of about 77,400 km2, much of which is conservation and protected land (48.78%) (Figure 3-15). A further 5.03% is classified as water and wetlands, most of which is in several large areas classified as marsh and wetlands (4.22%) throughout the northern parts of the Assessment area. Most of the remaining area (45.74%) is used for grazing natural vegetation. Intensive agriculture and cropping make up a very small portion of the catchment: dryland and irrigated agriculture and intensive animal production together comprise just 0.14% of the land area. The other intensive localised land uses are transport, communications, services, utilities and urban infrastructure (0.31%), and mining (less than 0.01% of the catchment area). While not considered a land use under the land use mapping (because it is a tenure) it is worth noting that Aboriginal freehold title, held under the Aboriginal Land Rights (Northern Territory) Act 1976 (Cth) makes up 45% of the Roper catchment. The title is inalienable freehold, which cannot be sold and is granted to Aboriginal Land Trusts which have the power to grant an interest over the land. Native title exists in parts of the native title determination areas that occur in an additional 37% of the catchment. Figure 3-15 Land use classification for the Roper catchment Source: Northern Territory Land Use Mapping Project 2016 to current, Department of Environment, Parks and Water Security, Northern Territory Government http://www.ntlis.nt.gov.au/metadata/export_data?type=html&metadata_id=5779F987695AE0FAE050CD9B21447ADC Se-R-514_Map_landuse_Roper_v3 For more information on this figure or equation please contact CSIRO on enquiries@csiro.au Value of agriculture and fisheries The value of agricultural production for the ABS SA4 region that covers the Roper catchment is given in Table 3-6, together with the estimated proportion of that production that occurs within the catchment. The value of agricultural production in the SA4 region was about 30% higher in 2019–20 than 2015–16, mainly due to increased gross revenue from the beef industry. The value of crops from the region has remained relatively stable over the same period (a 3% decline, Table 3-6). The most recent annual survey data from the ABS describing the value of agriculture by different types of industries (2019–20 survey), are only available at a much larger scale (SA4 level; see Figure 3-14) than the Roper catchment, making it difficult to accurately estimate the value of agriculture products within the catchment. Hence estimates have been made using 2015–16 agricultural census data, which were published by ABS at finer spatial scales (SA2 level), and then adjusting these by the ratio of the SA4 value for 2019–20 relative to 2015–16 (Table 3-6). Table 3-6 Value of agricultural production within the wider SA4 region and estimates of the value of agricultural production for the Roper catchment For more information on this figure or table please contact CSIRO on enquiries@csiro.au Estimate for Roper catchment based on SA2 data apportioned using weighted averages of scores for SA2 regions falling wholly or partially within the catchment boundary (ABS, 2017). 2019–20 estimate for Roper catchment based on applying percentage of ‘total crops’ and ‘total livestock’ by SA4 that fall within the catchment based on the ratio of 2019–20 data to the 2015–16 SA4 data (ABS, 2021). Sources: ABS (2017, 2021) Agriculture is a major source of employment in the Roper catchment, featuring within the top three industries by employment levels, as shown in Table 3-5. This is very different to the importance of agriculture to employment on a national basis. Beef cattle production Agricultural production in the Roper catchment is dominated by extensive grazing of beef cattle, valued at $55.5 million in 2019–20 (Table 3-6). The first cattle were brought to Elsey Station, near Mataranka, in 1882 (Gleeson and Richards, 1985). The life of European settlers at the Elsey Station and the surrounding region have become well known through the account given in the autobiographical novel We of the never-never (Gunn, 1908). Subsequent attempts at expanding livestock production along the Roper River initially met with mixed success. The Mataranka Horse and Sheep Experimental Station was established in 1913 as part of plans for closer settlement in the upper reaches of the Roper catchment, which envisioned agricultural-led growth in the region by which Mataranka would replace Darwin as the NT capital (Gleeson and Richards, 1985). However, sheep production proved unsuitable for the region because of the harsh climate, rough terrain and blowfly incidence, so was abandoned in favour of beef cattle, which proved more successful and persists as the dominant agricultural activity in the catchment to this day. The within-year variation produced by the wet-dry climate is the main determinant for cattle production. Native pasture growth is dependent on rainfall; therefore, pasture growth is highest during the December to March period. During the dry season, the total standing biomass and the nutritive value of the vegetation declines. Changes in cattle live weight closely follow this pattern, with higher growth rates over the wet season compared to the dry season. Indeed, in many cases cattle lose live weight and body condition throughout the dry season until the next pulse of growth initiated by wet-season rains. A whole-of-industry survey (Cowley, 2014) provides a snapshot of the industry as it was in 2010. While some of the survey results below have inevitably changed since then, the general enterprise type has not changed significantly in the last decade and the following can be considered still current. Cowley (2014) presents data for the whole of the Katherine region, broken into five districts: Roper, Sturt Plateau, Katherine/Daly, Victoria River and Gulf. The information below comes from either the Roper or the Sturt Plateau districts unless noted to be from the Katherine region as a whole (i.e. across all five districts). Note that the Roper and Sturt Plateau districts do not follow Roper catchment boundaries but can be considered broadly representative of those properties within the catchment. Further detail can be found in the companion technical report on agricultural viability and socio-economics (Stokes et al., 2023). The majority of properties in the Roper and Sturt Plateau districts were less than 1000 km2 in size, with a minimum of 20 km2 in the Roper. In both these districts, about 10% of properties were greater than 4000 km2. The average property size was 1133 km2 (Roper) and 1308 km2 (Sturt Plateau). Across the Katherine region as a whole, nearly 40% of properties were ‘Owner-Manager’. Typically, these are smaller enterprises with less cattle than ‘Company-Manager’ properties, with 48% of the cattle and 43% of the land under this latter category. Company-Manager properties are often part of an integrated enterprise, involving transfers of cattle between properties in the company and sharing of staff and resources. Owner-Manager properties are more likely to consist of only one property and run as a stand-alone enterprise. A large area of land is needed to maintain one unit of cattle (typically termed an AE, or adult equivalent). This carrying capacity of land is determined primarily by the soil (and landscape) type, the average annual rainfall and its seasonality, and the consequent native vegetation type. Northern Territory Government estimates of carrying capacity on the Sturt Plateau range from a maximum of 15 to 21 AE/km2 (i.e. 4.8 to 6.7 ha/AE) on the relic floodplains of the Larrimah land system in ‘A’ condition (from a four point scale where ‘A’ is highest and ‘D’ is lowest) to a low of 4.5 to 5.0 AE/km2 (i.e. 20 to 22.2 ha/AE) on ‘C’ condition pastures of the Elsey and Bulwaddy land systems, noting that ‘D’ condition lands across the region have a recommended carrying capacity of zero AE/km2 (Pettit, undated). Carrying capacity estimates for the alluvial plains of the Gulf Fall are not as readily available as for the Sturt Plateau but these types of landscapes are typically considered of ‘moderate to high’ or ‘high’ pastoral value. The typical beef production system is a cow-calf operation with sale animals turned-off for the live export market, via Darwin Port. The most common live export destination was South-East Asia. The most common grazing strategy is a combination of continuous grazing and wet-season spelling. Rotational grazing, or cell grazing, are not typically used. About 78% of all cattle across the Katherine region were Brahman, with about another 17% being Brahman derived. The majority of surveyed properties in both the Roper and Sturt Plateau districts ran between 2000 and 5000 head of cattle. Owner-Manager properties typically ran fewer cattle than Company-Manager properties. The majority of properties in both the Roper and Sturt Plateau districts breed cattle for the live export market, although a significant percentage (38% and 24%, respectively) bred cattle to transfer and grow-out elsewhere. Across the Katherine region, 83% of cattle turned-off made their way to live export either directly or indirectly through inter-company transfers, backgrounding or floodplain agistment, closer to Darwin. Across the Katherine region most of the cattle are sold off-property early in the dry season, at the time of the first round of mustering. The most common sales months are May to July, with a secondary peak in September–October. This corresponds to the common practice of two rounds of mustering, with the first early in the dry season and the second late in the dry season. While the cattle typically graze on native pastures, many properties supplementary feed hay to the weaner cohort, partly to train them to be comfortable around humans for management purposes and partly to add to their growth rates during the dry season when the nutritive value and total standing biomass of native pastures is falling. Urea-based supplements and supplements containing phosphorus are fed to a range of age and sex classes of the cattle. The urea-based supplements are to provide a source of nitrogen for cattle grazing dry-season vegetation while the phosphorus supplements, mostly provided over the wet season, are used because phosphorus is deficient in many areas yet it is required for many of the body’s functions such as building bones, metabolising food and producing milk (Jackson et al., 2015). Supplements were fed in 88% of the Roper district and 100% of the Sturt Plateau district. Cropping Cropping in the Roper catchment has an annual value of only about $18 million (Table 3-6), mainly from melons and mangoes (Mangifera indica) grown near Mataranka. Mataranka complements Katherine as a mango growing area since the climate is slightly cooler, which means that flowering and fruit ripening occur later and thus extends the overall duration of the harvest season for the region. Despite more than a century of attempts at establishing crop industries in the NT, there is still very little irrigated or dryland cropping in the Roper catchment. After the agricultural experiments around the time of the First World War, the Second World War prompted another wave of interest in facilitating northern agricultural development, which included a set of agricultural experimental stations. In 1942, approval was given to establish army farms at Katherine and Mataranka with the aim of more efficiently supplying the fruit and vegetables needed to maintain the nutrition of troops. The army experimental farm at Katherine was initially established to test what fruit and vegetables were suitable for the area. After the war this became the Katherine Experimental Station, where a wider range of crops were explored (run by the Commonwealth until it was handed over to the Northern Territory Government in the 1980s). Several crops, such as peanuts (Arachis hypogaea) in the 1950s, initially proved to be agronomically suitable for the local environment but were unable to be established as competitive local industries, partly because of difficulties with market access and high transport costs. Aquaculture and fisheries There is currently no active aquaculture in the Roper catchment. A freehold area of approximately 12,000 ha about 25 km upstream from the mouth of the Roper River was developed by Carpentaria Aquafarm Pty Ltd with about 40 ha of grow-out ponds in the 1980s. While the ponds remain, the business stopped operating in the early 1990s (Australasia Aquaculture, 2005). Offshore, the Roper River drains into one of the most valuable fisheries in the country. The NPF spans the northern Australian coast between Cape Londonderry in WA to Cape York in Queensland (Figure 3-16), with most of the catch being landed at the ports of Darwin, Karumba and Cairns. Over the 10-year period from 2010–11 to 2019–20, the annual value of the catch from the NPF has varied between $65 million and $124 million, with a mean of $100 million (Steven et al., 2021). The Roper catchment flows into the South Groote NPF region (Figure 3-16), one of the smallest regions by annual prawn catch. Like many tropical fisheries, the target species exhibit an inshore–offshore larval life cycle and are dependent on inshore habitats, including estuaries, during the postlarval and juvenile phases (Vance et al., 1998). Monsoon-driven freshwater flood flows cue juvenile prawns to emigrate from estuaries to the fishing grounds and flood magnitude explains 30 to 70% of annual catch variation, depending on catchment region (Buckworth et al., 2014; Vance et al., 2003). Fishing activity for banana and tiger prawns, which constitute 80% of the catch, is limited to two seasons: a shorter banana prawn season from April to June, and a longer tiger prawn season from August to November. The specific dates of each season are adjusted depending on catch rates. Banana prawns generally form the majority of the annual prawn catch by volume. Key target and by- product species are detailed by Woodhams et al. (2011). The catch is often frozen on-board and sold in domestic and export markets. Se-R-501_Portrait_map_Australia_NPF_regions_v3 For more information on this figure or equation please contact CSIRO on enquiries@csiro.au Figure 3-16 Map of regions in the Northern Prawn Fishery The regions in alphabetical order are Arnhem-Wessels (AW), Cobourg-Melville (CM), Fog Bay (FB), Joseph-Bonaparte Gulf (JB), Karumba (KA), Mitchell (ML), North Groote (NG), South Groote (SG), Vanderlins (VL), Weipa (WA), West- Mornington (WM). Source: Dambacher et al. (2015) The NPF is managed by the Australian Government (via the Australian Fisheries Management Authority) through input controls, such as gear restrictions (number of boats and nets, length of nets) and restricted entry. Initially comprising over 200 vessels in the late 1960s, the number of vessels in the NPF has reduced to 52 trawlers and 19 licensed operators after management initiatives including effort reductions and vessel buy-back programs (Dichmont et al., 2008). Given recent efforts to alleviate fishing pressure in the NPF, there is little opportunity for further expansion of the industry. However, any development of water resources in the Roper catchment would need to consider the downstream impacts on prawn breeding grounds and the NPF. 3.3.4 Current infrastructure Transport The most significant road in the Roper catchment is the Stuart Highway, which runs from Darwin to Port Augusta in SA, about 300 km north of Adelaide. The Stuart Highway is formally designated Route A1 from Darwin to Daly Waters and Route A87 from Daly Waters to Port Augusta. The road passes through Mataranka and Larrimah at the top of the catchment in the west and is the main link northwards to Katherine and Darwin and southwards to the south-eastern states via Alice Springs. Figure 3-17 shows the network of roads within the Roper catchment together with rankings according to the types of road surface. All road network information in this section is from spatial data layers in the Transport Network Strategic Investment Tool (TraNSIT: Higgins et al., 2015). Aside from the Stuart Highway, the Roper catchment is served by a sparse network of mainly unsealed roads. The most important roads branching off the Stuart Highway into the catchment are the Roper Highway (Route B20), linking Ngukurr near the mouth of the river (in the east) to Mataranka at the top of the catchment (in the west), and the Central Arnhem Road (Route C24), which runs across the north of the catchment from the Stuart Highway through Bulman/Gulin Gulin. Figure 3-18 shows the heavy vehicle access restrictions for roads within the Roper catchment, as determined by the National Heavy Vehicle Regulator. All non-residential roads in the study area permit Type 2 road trains, which are vehicles up to 53 m in length, typically a prime mover pulling three 40-foot trailers (Figure 3-19). Despite the poorer road conditions of many of the local unsealed roads, large (Type 2) road trains are permitted due to minimal safety issues from low traffic volumes and minimal road infrastructure restrictions (e.g. bridge limits, intersection turning safety). Drivers would regularly use smaller vehicle configurations on the minor roads due to the difficult terrain and single lane access, particularly during wet conditions. Figure 3-20 shows the Stuart Highway – the main north–south road corridor in the catchment. Figure 3-21 shows the speed limits for the road network within the Roper catchment. These speed limits are usually higher than the average speed achieved for freight vehicles, particularly on unsealed Rank 3 roads. Heavy vehicles using such unsealed roads would usually achieve average speeds of no more than 60 km/hour, and often as low as 20 km/hour when transporting livestock. A good quality standard gauge rail line passes through the western edge of the Roper catchment. This provides freight access to Darwin Port (East Arm Wharf) to the north, and to major southern markets via Alice Springs. The rail line is primarily used for bulk commodity transport (mostly minerals) to Darwin Port. There are no branch lines in the Roper catchment so goods would have to be transported to and from loading points by roads. Se-R-508_Roper_TraNSIT_road rankings_v2 For more information on this figure or equation please contact CSIRO on enquiries@csiro.au Figure 3-17 Road rankings and conditions for the Roper catchment Rank 1 = well-maintained highways or other major roads, usually sealed; Rank 2 = secondary ‘state’ roads; Rank 3 = minor routes, usually unsealed local roads. The ‘Rank 1’ road is the Stuart Highway, which runs from Darwin (in the north) to Port August (in SA, about 300 km north of Adelaide). Se-R-509_Roper_TraNSIT_truck_type_v2 For more information on this figure or equation please contact CSIRO on enquiries@csiro.au Figure 3-18 Vehicle access restrictions for the Roper catchment Truck classes referred to in the legend are illustrated in Figure 3-19. For more information on this figure, please contact CSIRO on enquiries@csiro.au Figure 3-19 Common configurations of heavy freight vehicles used for transporting agricultural goods in Australia For more information on this figure or equation, please contact CSIRO on enquiries@csiro.au Figure 3-20 Looking south along the Stuart Highway the main north–south transport artery of the Northern Territory Photo: CSIRO - Nathan Dyer Se-R-510_Roper_TraNSIT_road_speed_v3 For more information on this figure, please contact CSIRO on enquiries@csiro.au Figure 3-21 Road speed restrictions for the Roper catchment Supply chains and processing Table 3-7 provides volumes of agricultural commodities transported into and out of the Roper catchment and Figure 3-22 shows the location of existing agricultural enterprises in the catchment. As previously noted, agricultural production in the Roper catchment is currently dominated by horticulture and beef, particularly melons and live cattle export. This is also reflected in annual volumes of commodities transported into and out of the catchment through the road network. About 31,000 t of melons were transported out of the catchment, predominantly to domestic distribution centres in 2021, according to TraNSIT records of truck movements. There are also large volumes of freight transporting cattle into (~13,000 head in 2021) and out of (~36,000 head combined) the Roper catchment, mainly via the Stuart Highway. Live export of cattle via Darwin Port account for the majority of cattle movements, but there are also substantial transfers of cattle between properties and smaller volumes directed to domestic markets via abattoirs and feedlots. There are currently no processing facilities for agricultural produce within the Roper catchment, but there are (or soon will be) facilities nearby that could support producers in the catchment. The closest meatworks was run by AACo (Australian Agricultural Company) at Livingstone, about 40 km south of Darwin but has not been operational since 2018. When operating, it was accessible by large (Type 2) road trains along the entire route from the Roper catchment. The first cotton gin in the NT commenced construction in 2021 (scheduled for completion in 2023) about 30 km north of Katherine as part of growing interest in establishing a new cotton industry in the region. The closest port for bulk export of agricultural produce from the Roper catchment is in Darwin. Darwin Port, operated by Landbridge Group, handles about 20,000 to 30,000 twenty-foot equivalent units (TEU) each year split roughly evenly between imports and exports. The main exports are dry bulk commodities (mainly manganese) and livestock, but there are also annual exports of about 100 TEU of refrigerated containers. Exports of new bulk agricultural produce would require construction of a new storage facility. Table 3-7 Overview of agriculture commodities transported into and out of the Roper catchment Indicative transport costs are means for each commodity and include differences in distances between source and destinations. For more information on this figure or table please contact CSIRO on enquiries@csiro.au Source: 2021 data from TraNSIT (Higgins et al., 2015) Se-R-511_TraNSIT_Roper_ag enterprises_v4 For more information on this figure or equation please contact CSIRO on enquiries@csiro.au Figure 3-22 Agricultural enterprises in the Roper catchment and amount of annual trucking to/from them Smaller horticultural enterprises, mainly melon and mango farms near Mataranka, are too small to show on the map at this scale. The thickness of purple lines indicates how much traffic (as number of tailers) there is on regional roads connecting local enterprises. Energy The Darwin-Katherine Interconnected System (DKIS) is the largest electricity grid in the NT (Figure 3-23). The DKIS is electrically isolated from other grids in Australia (but see below for how electricity and natural gas transmission systems are interconnected). The DKIS transmission network reaches the western edge of the Roper catchment, passing through Mataranka and reaching as far south as Larrimah. A small branch off this main transmission line serves Barunga (Bamyili) and Beswick, and a distribution line links Jilkminggan to nearby Mataranka. Generation on the DKIS is primarily by gas turbine power stations at Channel Island (279 MW), Weddell (129 MW), Katherine (36.5 MW) and Pine Creek (26.9 MW, privately owned). The closest generator to the Roper catchment is at Katherine, where there is an additional back-up diesel generator. Most of the Roper catchment, however, is too remote to be covered by the DKIS. The three largest off-grid remote communities rely on hybrid systems powered by diesel generators supplemented with solar: Ngukurr (400 kW solar system), Minyerri (275 kW) and Bulman (100 kW). Distribution lines link nearby smaller settlements to these off-grid sources of electricity: Rittarangu is connected to Ngukurr and Weemol is connected to Bulman. Historically, gas pipelines have been a cheaper way of transporting energy than electrical transmission lines (DeSantis et al., 2021; GPA, 2021). So, a network of natural gas pipelines has been a cost-effective way of linking energy supplies across the NT by connecting sources of gas to electricity generators and other demand centres. The Amadeus Gas Pipeline is a bi-directional pipeline running from the gas fields of the Amadeus Basin near Alice Springs in the south northwards through the western edge of the Roper catchment (near the Stuart Highway) towards Darwin. The McArthur River Pipeline connects to the Amadeus Gas Pipeline at Daly Waters and runs across the southern edge of the Roper catchment to the generator at the McArthur River zinc-lead mine. The Northern Gas Pipeline, which runs 622 km between Tennant Creek and Mount Isa in Queensland, provides a connection between the energy systems of the NT and the eastern states. Se-R-507_Roper_energy_generation_distribution_v2 For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-23 Electricity generation and transmission network and natural gas pipelines in the Roper catchment Distribution networks are not shown, but communities marked with red lightning symbols are connected to nearby generation or transmission sources of electricity. The Amadeus Gas Pipeline runs north–south (bi-directional) through Katherine; the McArthur River Pipeline branches off eastwards from Daly Waters to the McArthur River zinc-lead mine. Inset shows pipeline and transmission network across the NT. Water The majority of the communities in the Roper catchment source their water from groundwater. Surface water is also used in some cases: river pumping supplements the water supply for Ngukurr, while outstations may source their domestic water requirements from river water, springs and lagoons (Zaar, 2009). There are no major dams or water transmission pipelines in the catchment. Surface water entitlements Licensed surface water entitlements are sparse across the Roper catchment which occupies the eastern part of the Daly Roper Beetaloo Water Control District (DRBWCD) The Northern Territory government prepares Water Allocation Plans to sustainably manage and allocate water resources, the Georgina Wiso Water Allocation Plan 2023 - 2031 (GWWAP) is currently in place with the Mataranka Tindall Limestone Aquifer Water Allocation Plan (MTLAWAP) (Figure 3-24) and the Surface Water Take – Wet Season Flows Policy currently being developed. There have been four licences granted for a combination of public water supplies and cultural and industrial uses, all of which fall outside of proposed water allocation planning areas. The largest entitlement of 80 ML/year is for public water supply at Barunga (Figure 3-24). The water is sourced from Beswick Creek which is fed by water discharging from Bamyili Spring. The next largest entitlement of 26 ML/year is from the Roper River for industrial purposes. Minor entitlements of <8 ML/year from the Roper River have been granted for cultural and industrial purposes. Groundwater entitlements Licensed groundwater entitlements of about 32.5 GL/year have been granted across the central and south-western parts of the Roper catchment, most prominently around Mataranka. Most of these entitlements are for water sourced from the regional-scale Cambrian Limestone Aquifer (CLA) in the south-west of the catchment. These licensed entitlements all occur within the proposed MTLAWAP with the exception of one licence for public water supply at Daly Waters (Figure 3-24). Only very minor entitlements (about 1 GL/year) are sourced elsewhere, mostly from localised fractured and weathered rock aquifers hosted in the Roper Group. The purpose of the majority of these entitlements is for irrigated agriculture (31.1 GL/year) all of which is sourced from the CLA. Licensed entitlements totalling about 0.4 GL/year have also been granted for public water supplies from the CLA for the communities of Mataranka, Jilkminggan, Larrimah and Daly Waters. The remainder of entitlements from the CLA (totalling about 0.3 GL/year) have been granted for industrial purposes including tourist accommodation, and council and cement operations. Other licensed entitlements come from aquifers hosting intermediate to local-scale groundwater systems inside the DRBWCD but outside of the proposed MTLAWAP and GWWAP areas (Figure 3-24). Licences have been granted for public water supply at Beswick (190 ML/year), Barunga (280 ML/year) and Minyerri (150 ML/year). Groundwater for Barunga is sourced from localised aquifers hosted in Cretaceous sandstone. Groundwater for Beswick is sourced from the intermediate-scale dolostone aquifer hosted in the Dook Creek Formation. Groundwater for Minyerri is sourced from a localised aquifer hosted in the Bessie Creek Sandstone. Groundwater is also used for stock and domestic water supplies for which a water licence is not needed. For more information on groundwater resources of the Roper catchment see the companion technical report on hydrogeological assessment by Taylor et al. (2023). For more information on this figure please contact CSIRO on enquiries@csiro.au Figure 3-24 Location, type and volume of annual licensed surface water and groundwater entitlements Data sources: Daly Roper Beetaloo Water Control District sourced from Department of Environment, Parks and Water Security (2019b); Water Allocation Plan areas sourced from Department of Environment, Parks and Water Security (2018) Community infrastructure The availability of community services and facilities in remote areas can play an important role in attracting or deterring people from living in those areas. Development of remote areas therefore also needs to consider whether housing, education and healthcare are sufficient to support the anticipated growth in population and demand, or to what extent these would need to be expanded. There are no hospitals in the Roper catchment but, like most remote parts of Australia, the area is serviced by a primary health network (PHN). Australia is divided into 31 PHNs and one of these covers the whole of the NT. General practitioners and allied health professionals provide most primary healthcare in Darwin and the regional centres within the Northern Territory PHN, while smaller communities are supported by remote health clinics (NT PHN, 2020). The Roper catchment falls within the Katherine Health Service District (HSD) (also known as the Big Rivers Region) of the Northern Territory PHN where the Sunrise Health Service Aboriginal Corporation and Katherine West Health Board provide remote health services. PHNs work closely with local hospital networks, and for the Katherine/Big Rivers Region the associated hospital is Katherine Hospital, which is located just outside the western border of the Roper catchment. This hospital has 60 beds and provides emergency services, surgical and medical care, paediatrics and obstetrics (NT PHN, 2020). A network of eight schools cover the small communities throughout the Roper catchment. A total of 807 full time equivalent (FTE) students are enrolled in these schools with 77.2 teachers (FTE) in 2021 (Table 3-8). The largest school in the catchment is at Ngukurr. There are a further six schools in Katherine, just outside the Roper catchment and about 100 km north-west of Mataranka, and there is also a school of the air in Katherine that serves 167.5 students (FTE) in the region. Table 3-8 Schools servicing the Roper catchment For more information on this figure or table please contact CSIRO on enquiries@csiro.au FTE = full time equivalent Source: ACARA (2022) (data presented with permission) At the time of the 2016 census, only about 11% of private dwellings were unoccupied, representing a similar proportion to the national average although slightly lower than the NT (Table 3-9). This suggests that the current pool of housing may have some capacity to absorb small future increases in population. Table 3-9 Number and percentage of unoccupied dwellings and population for the Roper catchment For more information on this figure or table please contact CSIRO on enquiries@csiro.au †Weighted averages of scores for SA2 regions falling wholly or partially within the catchment boundary. Source: ABS (2016) census data 3.4 Indigenous values, rights, interests and development goals 3.4.1 Introduction and research scope This section provides an overview of the existing information needs with respect to Indigenous water issues in the Assessment area to provide foundations for further community and government planning and decision making. Unless otherwise stated, the material in this section is based on findings described in the companion technical report on Indigenous aspirations, interests and water values (Lyons et al., 2023). Indigenous peoples represent a substantial and growing proportion of the population across northern Australia, and control significant natural and cultural resource assets, including land, water and coastlines. They will be crucial owners, partners, investors and stakeholders in future development. Understanding the past is important to understanding present circumstances and future possibilities. Section 3.4 provides some key background information about the Indigenous Australians of the Roper catchment and their specific values, rights, interests and objectives in relation to water and irrigated agricultural development. Section 3.4.2 describes the past habitation by Indigenous people, the significance of water in habitation patterns, and the impact of exploration and colonisation processes. Section 3.4.3 reviews the contemporary situation with respect to Indigenous residence, land ownership and access. Section 3.4.4 outlines Indigenous water values and responses to development, and Section 3.4.5 describes Indigenous-generated development objectives. The material provided here is a short summary of the research undertaken. Further details regarding this component of the Assessment are contained in the companion technical report on Indigenous aspirations, interests and water values (Lyons et al., 2023). There has been some previous information about Indigenous water values and historical water management in the upper reaches of the Roper catchment, with far less consideration of Indigenous perspectives on general water development and associated irrigated agricultural development in the region. The work undertaken here directly addresses these data needs. Engagement with Indigenous people is a strong aspiration across governments and key industries, but models of engagement can vary considerably and competing understandings of what ‘engagement’ means (consultation, involvement, partnership, etc.) can substantially affect successful outcomes. Standard stakeholder models can also marginalise Indigenous interests, reducing what Indigenous people understand as prior and inalienable ownership rights to a single ‘stake’, equivalent to all others at the table. Guided by advice from the Northern Land Council (NLC) and its Ngukurr Regional Council members, the Roper Gulf Regional Council Local Authority members and from senior Traditional Owners in the study area, the Assessment engaged nominated senior Indigenous leaders from within the Roper catchment in one-on-one and small group interviews to establish a range of views regarding water and agricultural development. The companion technical report on Indigenous aspirations, interests and water values (Lyons et al., 2023) provides details of these data and is a crucial supporting document for the summary provided here. A small set of comments are replicated in the following sections to show the type of data obtained, complemented by key themes analysed in the data. The Assessment does not provide formal Indigenous group positions about any of the issues raised and does not substitute for formal processes required by cultural heritage, environmental impact assessment, water planning, or other government legislation. Nevertheless, the research undertaken for this component of the Assessment identifies key principles, important issues and potential pathways to provide effective guidance for future planning and for formal negotiations with Indigenous groups. 3.4.2 Pre-colonial and colonial history Pre-colonial Indigenous society Pre-colonial Indigenous society is distinguished by four primary characteristics: long residence times; detailed knowledge of ecology and food gathering techniques; complex systems of kinship and territorial organisation; and a sophisticated set of religious beliefs, often known as Dreaming. These Indigenous religious cosmologies provided a source of spiritual and emotional connection as well as guidance on identity, language, law, territorial boundaries and economic relationships (Merlan, 1982; Rose, 2002; Strang, 1997; Williams, 1986). From an Indigenous perspective, ancestral powers are present in the landscape in an ongoing way, intimately connected to people, country and culture. Mythological creators, referred collectively as Dreaming, have imbued significance to places through creation, leaving evidence of their actions and presence through features in the landscape (Merlan, 1981, 1982). The cosmological belief of Dreaming is present among many Indigenous groups. Totemic figures can be animals or plants, take human-like or inanimate object form, or be sentient beings that have agency to act (Merlan, 1982; Peterson, 2013). Those powers must be considered in any action that takes place on the country. Northern Australia contains archaeological evidence of Indigenous habitation stretching back many thousands of years (Clarkson et al., 2017), but there remain gaps in the published archaeological record. Resource-rich riverine habitats were central to Indigenous economies based on seasonally organised hunting, gathering and fishing. Rivers were also major corridors for social interaction, containing many sites of cultural importance (Barber and Jackson, 2014; McIntyre-Tamwoy et al., 2013). Colonisation European colonisation resulted in significant levels of violence towards Indigenous Australians, with consequent negative effects on the structure and function of existing Indigenous societies across the continent. Avoidance, armed defensiveness and overt violence were all evident in colonial relationships as hostilities occurred as a result of competition for food and water resources, colonial attitudes and cultural misunderstandings. In the Roper catchment, the events that had the most significant impact on Indigenous people of the region were (Morphy and Morphy, 1981): • development of a supply depot for the Overland Telegraph on the Roper River • establishment of a stock route from Queensland to the north-west of the NT (through Borroloola and the Roper catchment) • gold rushes in Pine Creek and Kimberly in the 1880s that brought many miners through the region • establishment of permanent cattle stations. The Roper catchment was opened for colonial industry by the South Australian Government after John McDouall Stuart’s 1862 exploratory expedition reported plentiful water, fertile soils and native vegetation (Merlan, 1978, 1986). Pursuing expansion of its pastoral country, the South Australian Government physically annexed the NT in 1863 (Merlan, 1978; Zoellnner, 2017). A team of surveyors was sent to the Roper River in 1870 to assess the suitability of the river as a port for supplies. The Roper Landing (Roper Bar) was established as a supply depot for the telegraph parties. The construction and the establishment of the Overland Telegraph Line, which was completed in 1872, facilitated the first incursion that made an impact on the tribes of the Roper River (Merlan, 1978). The construction of the Overland Telegraph Line initiated intensive contact between Europeans and the local Indigenous peoples. Surveying and work force expedition encounters with Indigenous peoples were often violent and involved a lack of knowledge and misunderstanding about each of the others’ intentions and interests (Merlan, 1978). For the following three decades the government sought to establish a permanent European presence in the region. The first pastoral lease application at the top of the catchment was made in 1877 on an area that became part of Elsey station (Merlan, 1978). The station was stocked with cattle driven from NSW, through Roper Bar, Mount McMinn, Mole Hill and the Strangways River, and set up temporary camp at Crescent Lagoon. In the final advancement, cattle were moved through Red Lilly Lagoon onto Elsey Creek where they were released (Merlan, 1978). Pastoral occupation began in the early 1880s and was a focus for conflict as pastoral homesteads and outstations were sited close to permanent water and on the fertile plains and river valleys used by Indigenous people for food and other resources (McGrath, 1987; Merlan, 1986). Indigenous attacks on colonial pastoral operations were made both in retaliation for past attacks by colonists and as a response to shortages of food and other resources. In response, pastoralists responded with punitive expeditions, gaining greater influence as the Indigenous guerrilla war tactics became less effective with the expansion of pastoralism into new country (Merlan, 1982; Sandefur, 1985). Figure 3-25 shows the general areas of colonial violent encounters in the Roper catchment. CR-R-Ch3_500_Roper_Rivers_Massacres_Indigenous For more information on this figure or equation or table, please contact CSIRO on enquiries@csiro.au Figure 3-25 Colonial frontier massacres in the Roper catchment Source: Ryan et al. (2018), also see https://c21ch.newcastle.edu.au/colonialmassacres/ (accessed 15 March 2023). Local group structures were severely disrupted from the early 1880s, at which time Traditional Custodians established themselves at cattle stations and mission settlements (Merlan, 1981). Stations became places for enforced dependence and colonial influence to control Indigenous peoples, protect cattle and potentially incorporate them as assets to the development vision of northern Australia (Merlan, 1978, 1982). There was consequential reduction in Indigenous numbers in the early phase of European settlement from violence and disease, though not all encounters were violent (Merlan, 1986). The socio-linguistic groups that identified with the Roper region largely remained in the area on stations and settlements (Merlan, 1981). In the areas of Elsey and Roper Valley stations, the local Indigenous people have sustained their relationships to lands through their association within localised ‘countries’ (Merlan, 1986). Both Elsey and Hodgson Downs stations were bought by the Eastern and African Cold Storage Co. Ltd in 1903. The intention of the Eastern and African Cold Storage Co. was to stock its 20,000 square miles of leased coastal frontage in the Blue Mud Bay region by moving cattle from the Elsey-Hodgson region, what was then thought to be rich pasture country on the rivers of the Roper catchment. As Merlan (1978, p. 87) writes ‘[i]n the six years of its operation the ‘Eastern and African’ engaged in what was apparently the most systematic extermination of Aborigines ever carried out on the Roper’. 3.4.3 Contemporary Indigenous ownership, management, residence and representation Despite the pressures entailed by colonisation, country remained crucial to Indigenous peoples’ lives, sustaining a distinct individual and group identity as well as connections to past ancestors and future descendants. People are connected to places through a combination of genealogical, traditional and residential ties. Only some of these connections are formally recognised by the Australian state. Indigenous ownership The Indigenous owners of the Roper catchment include the Wubulawan, Bagala, Dalabon, Mangarrayi, Ngalakan, Ngandi, Warndarrang and Alawa peoples. There are also a range of related groups and subgroups within these regional ownership descriptors. The Assessment focused on the upstream and remote areas of the Roper catchment. Ownership patterns tend to follow natural landscape features such as rivers and hills, as well as formal boundaries between ownership groups, where these been negotiated. However, in other places the edges of group territory are less distinct and/or there are overlapping claims. Information regarding the identification of potential owners and interest holders is provided by registered organisations such as the NLC and the Aboriginal Areas Protection Authority (AAPA). In the NT jurisdiction there is specific land rights legislation that covers a wide area of the NT, namely the Aboriginal Land Rights (Northern Territory) Act 1976 (Cth) (ALRA). This provides a form of collective freehold ownership that is significant across the Roper catchment and includes about over 45% of the land area (Figure 3-26). CR-R-Ch3_501_Roper_Indigenous_freehold_ILUA For more information on this figure or equation or table, please contact CSIRO on enquiries@csiro.au Figure 3-26 Indigenous freehold (Aboriginal Land) in the Roper catchment as at July 2017 ILUA = Indigenous Land Use Agreement; ALRA = Aboriginal Land Rights (Northern Territory) Act 1976 (Cth) Data source: Northern Land Council Across the whole of Australia, the primary form of recognition for Indigenous interests is native title and associated Indigenous Land Use Agreements (ILUAs). Native title provides a series of rights (such as access, hunting and fishing) determined through a legal process. ILUAs are voluntary registered agreements between native title claimants or holders and other interested parties for the use and management of land and resources. There are two ILUAs in the Roper catchment, one in the township of Urapunga and another in Mataranka, covering less than 0.001% of the catchment area. Indigenous native title interests are currently formally recognised by the Australian legal system and exist in parts of determination areas that cover 37% of the Roper catchment. Figure 3-27 shows native title claims and determinations, and ILUAs. CR-R-Ch3_502_Roper_NativeTitle For more information on this figure or equation or table, please contact CSIRO on enquiries@csiro.au Figure 3-27 Indigenous native title claims and determinations in the Roper catchment as at July 2017 Data sources: National Native Title Tribunal and Northern Land Council Indigenous population and residence The Indigenous population as a percentage of the total, comprise 73.4% in the Roper catchment, as at 2016 (Table 3-3). This includes Indigenous people who are part of the recognised local ownership groups identified above, as well as residents who identify as Indigenous but have their origins elsewhere. For many Traditional Owners, primary residential locations may be outside of the traditional lands to which they have formal ties. These patterns of residence and dispersal reflect a combination of historical involuntary relocation, voluntary movement to seek jobs and other opportunities, and kinship and family links. As such, these administrative counts do not account for the complexity of Indigenous peoples’ social, linguistic and economic relations. Indigenous communities in the Roper catchment face a range of social and demographic challenges, including significant unemployment, poor health and housing, water insecurity, structural impediments to economic participation including remoteness and social and family units under high levels of stress. Assessment participants sought economic and social conditions that would enable more of their people, particularly the youth, to be employed, and for capacity to engage in formal planning processes on their own traditional lands, as two responses to these circumstances. Indigenous governance and representation Indigenous organisational and political structure within the Roper catchment is diverse. The NLC is the major regional Indigenous representative organisation for the Roper catchment, representing and acting for Traditional Owners with respect to Indigenous access, participation, partnership and ownership. Local groups in the area are represented through a range of Indigenous corporations and entities including Aboriginal land trusts. Amongst the Indigenous groups consulted for this Assessment, there were variations in existing capacity, resourcing, partnerships and ability to participate in natural resource management decision making. There was strong interest and support for the involvement of local leaders and the youth in natural resource management negotiation and planning processes. 3.4.4 Indigenous water values and responses to development Introduction: attachment, ownership, protection Indigenous values in relation to their country in the Roper catchment encompass principles of attachment, ownership, and the responsibility to protect and sustain country and connections. These are manifested in practical terms through: • the assumption of Indigenous ownership of land and water resources • the need for formal external recognition of and engagement with that ownership and its associated responsibilities • the role of local histories in establishing local Indigenous connections and authority • the role of rangers in land and sea management • the ongoing role of religious and spiritual beliefs • the existence of ongoing knowledge and practices that sustain group and language boundaries and identities • the importance of hunting, foraging and fishing activity to Indigenous cultures • sustained connections and livelihoods on outstations • inter-generational obligations to both ancestors and descendants to care for the country • regional responsibilities to near neighbours and downstream groups to maintain the integrity of the country and related customary knowledge and practices. These principles also apply to Indigenous attitudes to non-Indigenous activities on Indigenous lands. Four frequently highlighted principles are: • consultation with the relevant owners and impacted groups • consent for development • compliance with the terms of policies and agreements, including Traditional Owner employment and capacity building • compensation for the access to and use of resources. These principles have clear implications for native title, cultural heritage and environmental impact assessment, as well as for broader issues of sustainable development. Cultural heritage Indigenous cultural heritage is a crucial manifestation of the principles of attachment, ownership and protection. Cultural heritage itself has a number of components: archaeological sites; places associated with traditional stories or traditional knowledge; and places of historical or contemporary importance. Cultural heritage is strongly correlated with permanent water, meaning that riverine and aquatic areas that are the focus of development interest are also likely to contain significant cultural heritage. Traditional Owners expressed their strong preference for open access to cultural sites but also recognised the lease rights held by others and accordingly some have negotiated access arrangements with lease holders. Consultations between development proponents and Traditional Owners will be significantly aided by early stage field scoping of cultural heritage issues and requirements. Contemporary Indigenous water values In general terms, Indigenous water values emphasise securing sufficient water of good quality to maintain healthy landscapes, remote community health and livelihoods, and to support Indigenous needs. Those needs can be defined in multiple ways, and from an economic perspective encompass such activities as art and cultural production, hunting and gathering as well as pastoralism, water-related revenues, agriculture, educational and cultural tourism and fishing permits. All of these needs depend on natural resources, highlighting the importance of securing and maintaining good quality water supplies. Data from the Assessment clearly demonstrate the overall importance of water for community life, health and practical hygiene, religious symbolism and Dreaming, ancestral connection and the challenges Traditional Owners can experience in observing and compromising their values under continued development activities and interests: Water is life. Can’t keep saying yes to money all the time. Water is important for connections between groups. Water connects us here to Minyerri, all the way up the top. We’re connected through stories. We go camping, fishing, and hunting. Water is alive, everybody knows that. Traditional Owner and Minyerri Resident 1. Statements about the importance of water from participants in the Assessment are consistent with broader statements that outline significant Indigenous water rights, values and interests, both in Australia (NAILSMA, 2008, 2009) and internationally (United Nations, 2023; World Water Council, 2003). Responses to water and irrigation development In the Roper catchment, Indigenous responses to water and irrigation development are interpreted through perceptions of past and current development within and beyond the catchment, and through observations of ongoing environmental and seasonal changes. Indigenous responses to water development and extraction included considerations of impacts on water quality, on streamflow and on water-dependent ecosystems, community water access, and human cultural practices and recreation. Large instream dams were not favoured, and in general, larger scale water and agricultural development was seen as incompatible with contemporary Indigenous values and ways of living. Indigenous concerns about water development encompassed concerns about the cumulative impacts from other industries, particularly mining. Awareness of their position as long-term custodians, their marginalised socio-economic status, limited understanding of water planning processes and the ongoing impacts of current development projects, affected Indigenous peoples’ assessments of the relative risks and benefits associated with development proposals. Noting the above cautions, Indigenous participants also recognised that power imbalances may see large-scale development proceed. In this context, some data on preferences for particular kinds of water development were gathered, and the general trend from most to least favourable was: 1. Flood harvesting to supply smaller, offstream storages 2. Bore and groundwater extraction in the upper and mid reaches of the catchment 3. Smaller instream dams constructed inside tributaries or branches, which do not restrict all of the flow, across the catchment 4. Large instream dams in major river channels. Proposals for specific sites may not accord with this general trend, and new information may alter the above order at both local and regional scales. With respect to major water and irrigation development, key Indigenous criteria for evaluation include: • early and further formal consultations with Traditional Owners and affected groups, about options, environmental assessments and potential impacts and preferences • development that specifically addresses Indigenous needs (for example, education, amenity, access to sites, community and outstation water supply, and recreational opportunities) • appropriate cultural heritage surveys of likely areas of impact • agreements that support Indigenous employment and other benefits, and continuous consultation and assessment during development, construction and operation • the need for ongoing monitoring and reporting of resource use and its impacts that involves the employment of Traditional Owners • support for Indigenous roles in development projects that connect water development with water planning, wider catchment management and enterprise development. Indigenous interests in water planning Water planning is understood as one way of managing water development risk, but water planning also has particular challenges. In Australia, the National Water Initiative set the goal that jurisdictionally based water plans need to recognise Indigenous needs in terms of access and management (Department of Agriculture and Water Resources, 2017). This encompasses Indigenous representation, incorporation of Indigenous social, spiritual and customary objectives, and recognition of native title needs and uses. However, progress in implementing that recognition has been slow due to a lack of knowledge about those interests, competing water demands and the challenges of accommodating Indigenous perspectives in conventional planning frameworks. However, a new water planning initiative has been publicised for Indigenous landowners in the NT. Known as the Strategic Aboriginal Water Reserve (Northern Territory Government, 2017), this policy provides scope for further Indigenous recognition by creating reserved water allocations in water allocations plans for Indigenous development purposes. A water allocation plan for the Mataranka Tindal Limestone Aquifer is yet to be declared. Based on the data generated during the Assessment, formalising and refining Indigenous water values and water planning issues in the Roper catchment may require: • formal scoping discussions at local and catchment scales about how best to support Indigenous peoples’ involvement in water planning • refinement of Indigenous governance rights, roles and responsibilities in water planning • resourcing of Indigenous involvement in water planning, including formal training and water literacy programs • Indigenous-specific water allocations for development purposes that may include options for leasing water rights, and remote community and outstation water access and supply • further specification of the impacts on current and potential future native title rights and on cultural heritage • articulating water planning processes with land and development planning • addressing continuing Indigenous water research needs and information priorities. These propositions are based on the condition that Traditional Owners have relevant information for their decision making and have sufficient time to undertake their consultations at local and catchment scales. The Assessment highlighted the importance of literacy programs that improve understanding of: • water policy – water rights and entitlements, water for the environment and water for development and community livelihood • water plans – water allocations, access and volumes of allocation, Indigenous involvement in water planning, monitoring and management (see aforementioned set of principles), current and projected water uses • catchment hydrology – types of water sources for consumptive and non-consumptive uses (community, outstations, development projects), volumes and reliability of water supply, sensitivity of water sources and their quality and supply to climate change and projected uses. Water literacy programs for current and emerging leaders is perceived critical to building confidence for Indigenous peoples to engage actively in water planning and to promote and secure their current and future rights, interests and values in water. 3.4.5 Indigenous development objectives Indigenous people have a strong desire to be understood as development partners and investors in their own right and have generated their own development objectives. This stance informs responses to development proposals outlined by others. As a group, Indigenous people are socially and economically disadvantaged, but also custodians of ancient landscapes. They therefore seek to balance short- to medium-term social and economic needs with long-term cultural, historical and religious responsibilities to ancestral lands. Past forums have outlined Indigenous development agendas that are consistent with Indigenous perspectives in the Roper catchment (NAILSMA, 2012, 2013). These agendas are informed by two primary goals: • greater ownership of and/or management control over traditional land and waters • sustainable retention and/or resettlement of Indigenous people on their country. These goals are interrelated, because retention and/or resettlement relies on employment and income generation, and the majority of business opportunities identified by Indigenous people are land and natural-resource dependent: pastoralism, conservation services, ecotourism, agriculture, aquaculture and marine harvesting permits. Each group in the Roper catchment has multiple responsibilities and management roles but, based on geography, accessibility, residence, assets, governance and/or skills, some may more easily be able to sustain multiple business activities, while others may achieve greater success focusing on a single activity. Partnerships and planning Indigenous people in the Roper catchment possess valuable natural, historical and cultural assets and represent a significant potential labour force, but collectively lack business development skills and expertise. Partnerships can address this gap, such as those being facilitated by Centrefarm with the Mangarrayi and Wubalawan Land Trusts, but there remains a need to improve the opportunities for business to understand and invest in Indigenous people and lands in the Roper catchment. The development of a full business analysis may include the following actions: • investigation of the full range of potential business activities and options • production of group and/or catchment plans and prospectuses to coordinate and define collective Indigenous assets and opportunities and to aid communication with potential investors • further information and training for Indigenous people about the opportunities and constraints of partnerships with private industry, including effective use of Indigenous resource rights (land ownership and lease-holding native title, future water allocations, etc.) • targeted non-Indigenous community training regarding partnerships with Indigenous people, including models for shared benefit agreements and partnership arrangements, employment and training opportunities, etc. • creation of incentives for Indigenous involvement in new development initiatives, including relocation and resettlement allowances, pathways from training to jobs, employer incentives to hire and retain Indigenous staff, etc. • training for younger Indigenous people about career planning as well as formal job skills. Indigenous development objectives, and Indigenous development partnerships, are best progressed through locally specific, group and community-based planning and prioritisation processes that are nested in a system of regional coordination. Such planning and coordination can greatly increase the success of business development and of the opportunities for Indigenous employment, retention and resettlement that arise from them. Significant returns on investment may be achievable through well-targeted resourcing to local Indigenous entities, particularly Land Trusts and Aboriginal Corporations, to build understanding of business priorities and development objectives, as well as regional coordination processes, such as water planning and catchment management. Beyond business conditions, health and community services and infrastructure will attract and retain a skilled labour force. 3.5 Legal and policy environment Water planning Water plans include rules for managing licences and permits, including water trading rules if applicable. The Daly Roper Beetaloo Water Control District (DRWCD) includes the Roper catchment, and there is one water allocation plan (WAP), the Mataranka WAP (under development), within the Roper catchment. Aboriginal Water Reserve The NT legally requires that the allocation of water for Aboriginal use is part of water planning. The Strategic Aboriginal Water Reserve (SAWR) became statue in the NT in 2019. The SAWR is ‘water allocated in a WAP for Aboriginal economic development in respect of eligible land’ (Section 4(1), Water Act 1992 (NT)). At its maximum, the SAWR can be no more than 30% in an area with more than 30% of eligible Aboriginal land (Godden et al., 2020). An Aboriginal Water Reserve can only exist where there is eligible land at the time of the WAP. The draft Mataranka WAP provides for a notional SAWR. It includes the estate of four Traditional Owner groups: Bagala, Yangman, Mangarrayi and Wubulwun who own 53% of the land within the draft WAP area as Aboriginal freehold under the Aboriginal Land Rights (Northern Territory) Act 1976 (Cth) (Nikolakis and Grafton, 2022). Over 80% of the land in the area is categorised as eligible Aboriginal land for water planning processes (O’Donnell et al., 2022). The availability of water for the SAWR in the draft Mataranka WAP is uncertain under existing agreed levels of water allocations (Northern Territory Government ,2019, in O’Donnell et al., 2022). 3.6 References ABS (2006) Census of Population and Housing time series profile. Catalogue number 2003.0 for various SA regions falling partly within Roper catchment. Australian Bureau of Statistics, Canberra. Viewed 19 October 2021, https://www.abs.gov.au/census. ABS (2011) Census of Population and Housing time series profile. Catalogue number 2003.0 for various SA regions falling partly within Roper catchment. Australian Bureau of Statistics, Canberra. Viewed 19 October 2021, https://www.abs.gov.au/census. ABS (2016) Census of Population and Housing time series profile. Catalogue number 2003.0 for various SA regions falling partly within Roper catchment. Australian Bureau of Statistics, Canberra. Viewed 19 October 2021, https://www.abs.gov.au/census. ABS (2017) 75030DO005_201516 Value of agricultural commodities produced, Australia – 2015- 16, datacube released 31/10/17. Australian Bureau of Statistics, Canberra. Viewed 11 November 2021, https://www.abs.gov.au/statistics/industry/agriculture/value-agricultural- commodities-produced-australia. ABS (2021) 75030DO001_201920 Value of agricultural commodities produced, Australia – 2019- 20, datacube released 14/05/2021. Australian Bureau of Statistics, Canberra. Viewed 11 November 2021, https://www.abs.gov.au/statistics/industry/agriculture/value-agricultural- commodities-produced-australia. ACARA (2022) MySchool web site. Australian Curriculum, Assessment and Reporting Authority, Sydney. Viewed 13 July 2022, https://www.myschool.edu.au. AFMA (2022) Northern Prawn Fishery Harvest Strategy 2022. Australian Fisheries Management Authority, Canberra. ALA (2021) Atlas of Living Australia: Roper River Catchment occurrence. Viewed 22 November 2021, https://biocache.ala.org.au/occurrences/search?q=qid:1637538501985. ALA (2022) Atlas of Living Australia. Viewed 01 June 2022, http://www.ala.org.au. Alberts-Hubatsch H, Lee SY, Diele K, Wolff M and Nordhaus I (2014) Microhabitat use of early benthic stage mud crabs, Scylla serrata, in Eastern Australia. Journal of Crustacean Biology 34(5), 604–610. Alberts-Hubatsch H, Lee SY, Meynecke J-O, Diele K, Nordhaus I and Wolff M (2016) Life-history, movement, and habitat use of Scylla serrata (Decapoda, Portunidae): current knowledge and future challenges. Hydrobiologia 763(1), 5–21. Arthur A, Reid J, Kingsford R, McGinness H, Ward K and Harper M (2012) Breeding flow thresholds of colonial breeding waterbirds in the Murray-Darling Basin, Australia. Wetlands 32(2), 257– 265. DOI: 10.1007/s13157-011-0235-y. Australasia Aquaculture (2005) Low intensity prawn farming. Australasia Aquaculture (March 2005): 30–36. Australian Government (1999) Environment Protection and Biodiversity Conservation Act 1999. Australian Government, Canberra. Barber M and Jackson S (2012) Indigenous water management and water planning in the upper Roper River, Northern Territory: history and implications for contemporary water planning. CSIRO, Darwin. Barber M and Jackson S (2014) Autonomy and the intercultural: interpreting the history of Australian Aboriginal water management in the Roper River catchment, Northern Territory. Journal of the Royal Anthropological Institute 20(4), 670–693. DOI: 10.1111/1467- 9655.12129. Bayliss P, Buckworth R and Dichmont C (2014) Assessing the water needs of fisheries and ecological values in the Gulf of Carpentaria. Final report prepared for the Queensland Department of Natural Resources and Mines. CSIRO, Australia. Bino G, Steinfeld C and Kingsford RT (2014) Maximizing colonial waterbirds’ breeding events using identified ecological thresholds and environmental flow management. Ecological Applications 24(1), 142–157. DOI: 10.1890/13-0202.1. BirdLife International (2022) Important bird areas factsheet: Limmen Bight. Viewed 30 March 2022, http://datazone.birdlife.org/site/factsheet/limmen-bight-iba-australia/text. Bishop K, Allen S, Pollard D and Cook M (1990) Ecological studies on the freshwater fishes of the Alligator Rivers Region, Northern Territory: autecology. Supervising Scientist Report 145. Supervising Scientist, Darwin. Böhm M, Collen B, Baillie JEM, Bowles P, Chanson J, Cox N et al (2013) The conservation status of the world’s reptiles. Biological Conservation 157, 372–385. DOI: 10.1016/j.biocon.2012.07.015. Bowman DMJS (2000) Australian rainforests: islands of green in a land of fire. Cambridge University Press, Cambridge, New York. Bradley J (2018) Marra Sea Country: a report detailing the Limmen Bight Sea Country and cultural values. Department of Tourism Sport and Culture, Northern Territory Government, and Northern Land Council, Darwin. Broadley A, Stewart‐Koster B, Kenyon RA, Burford MA and Brown CJ (2020) Impact of water development on river flows and the catch of a commercial marine fishery. Ecosphere 11(7), e03194. Brodie JE and Mitchell AW (2005) Nutrients in Australian tropical rivers: changes with agricultural development and implications for receiving environments. Marine and Freshwater Research 56(3), 279–302. DOI: 10.1071/MF04081. Buckworth RC, Venables WN, Lawrence E, Kompas T, Pascoe S, Chu L, Hill F, Hutton T and Rothlisberg PC (2014) Incorporation of predictive models of banana prawn catch for MEY- based harvest strategy development for the Northern Prawn Fishery. Final report to the Fisheries Research and Development Corporation, Project 2011/239. CSIRO Marine and Atmospheric Research, Brisbane. Bunn SE and Arthington AH (2002) Basic principles and ecological consequences of altered flow regimes for aquatic biodiversity. Environmental Management 30, 492–507. Canham CA, Duvert C, Beesley LS, Douglas MM, Setterfield SA, Freestone FL, Clohessy S and Loomes RC (2021) The use of regional and alluvial groundwater by riparian trees in the wet‐ dry tropics of northern Australia. Hydrological Processes 35(5), e14180. Cann J and Sadlier R (2017) Freshwater turtles of Australia. CSIRO, Melbourne. Chan TU, Hart BT, Kennard MJ, Pusey BJ, Shenton W, Douglas MM, Valentine E and Patel S (2012) Bayesian network models for environmental flow decision making in the Daly River, Northern Territory, Australia. River Research and Applications 28(3), 283–301. Chatto R (2006) The distribution and status of waterbirds around the coast and coastal wetlands of the Northern Territory. Technical report. 76. Parks and Wildlife Commission of the Northern Territory, Darwin. Clarkson C, Jacobs Z, Marwick B, Fullagar R, Wallis L, Smith MA, Roberts RG, Hayes E, Lowe K, Carah X, Florin SA, McNeil J, Cox D, Arnold LJ, Hua Q, Huntley J, Brand HEA, Manne T, Fairbairn A, Shulmeister J, Lyle L, Salinas M, Page M, Connell K, Park G, Norman K, Murphy T and Pardoe C (2017) Human occupation of northern Australia by 65,000 years ago. Nature 547, 306–310. DOI: 10.1038/nature22968. Close P, Wallace J, Bayliss P, Bartolo R, Burrows D, Pusey B, Robinson C, McJannet D, Karim F and Byrne G (2012) Assessment of the likely impacts of development and climate change on aquatic ecological assets in Northern Australia. A report for the National Water Commission, Australia. Tropical Rivers and Coastal Knowledge (TraCK) Commonwealth Environmental Research Facility. Charles Darwin University, Darwin. Close P, Bartolo R, Pettit N, Ward D and Kennard M (2015) Vulnerability and risk assessment of northern Australian catchments and biodiversity. Climate Change Adaptation across Australia’s Monsoonal North–Northern Monsoon NRM Cluster, Griffith University, Nathan. Costermans LF (1981) Native trees and shrubs of south-eastern Australia. Cowley T (2014) The pastoral industry survey – Katherine region. Northern Territory Government, Australia. Craig LS, Olden JD, Arthington AH, Entrekin S, Hawkins CP, Kelly JJ, Kennedy TA, Maitland BM, Rosi EJ, Roy AH, Strayer DL, Tank JL, West AO and Wooten MS (2017) Meeting the challenge of interacting threats in freshwater ecosystems: a call to scientists and managers. Elementa: Science of the Anthropocene 5(72). https://doi.org/10.1525/elementa.256. Crocos PJ and Kerr J (1983) Maturation and spawning of the banana prawn Penaeus merguiensis de Man (Crustacea: Penaeidae) in the Gulf of Carpentaria, Australia. Journal of Experimental Marine Biology and Ecology 69(1), 37–59. Crook DA, Lowe WH, Allendorf FW, Erős T, Finn DS, Gillanders BM, Hadweng WL, Harrod C, Hermoso V, Jennings S, Kilada RW, Nagelkerken I, Hansen MM, Page TJ, Riginos C, Fry B and Hughes JM (2015) Human effects on ecological connectivity in aquatic ecosystems: integrating scientific approaches to support management and mitigation. Science in the Total Environment 534(2015), 52–64. Crook D, Buckle D, Allsop Q, Baldwin W, Saunders T, Kyne P, Woodhead J, Maas R, Roberts B and Douglas M (2016) Use of otolith chemistry and acoustic telemetry to elucidate migratory contingents in barramundi Lates calcarifer. Marine and Freshwater Research 68(8), 1554– 1566. Crook DA, Buckle DJ, Morrongiello JR, Allsop QA, Baldwin W, Saunders TM and Douglas MM (2019) Tracking the resource pulse: movement responses of fish to dynamic floodplain habitat in a tropical river. Journal of Animal Ecology 89(3), 795–807. Crowley G (2010) Management guidelines for the threatened species of the Northern Territory. Tropical Savannas Cooperative Research Centre, Darwin. Dambacher JM, Rothlisberg PC and Loneragan NR (2015) Qualitative mathematical models to support ecosystem-based management of Australia’s Northern Prawn Fishery. Ecological Applications 25(1), 278–298. Daniels CW, Ngukurr Yangbala Rangers, Russell S and Ens EJ (2022) Empowering young Aboriginal women to care for Country: case study of the Ngukurr Yangbala rangers, remote northern Australia. Ecological Management & Restoration 23(S1), 53–63. https://doi.org/10.1111/emr.12538. Dasgupta S, Blankenspoor B, Herzog T and Davies A (2019) Terrestrial biodiversity indicators. Viewed 10 October 2019, https://datacatalog.worldbank.org/dataset/terrestrial- biodiversity-indicators. David B, Taçon PSC, Delannoy JJ and Geneste JM (2017) The archaeology of rock art in Western Arnhem Land, Australia. Terra Australis, ANU Press, Canberra. Davies P, Bunn S and Hamilton S (2008) Primary production in tropical streams and rivers. In: Dudgeon D (ed.) Tropical stream ecology. Academic Press, Elsevier Inc, San Diego, California. Delaney R (2012) Limmen Bight: marine and coastal biodiversity values. Technical report. Northern Territory Department of Parks and Wildlife Service, Darwin. https://depws.nt.gov.au/__data/assets/pdf_file/0009/255087/Limmen-Marine-Technical- Report.pdf. Department of Agriculture and Water Resources (2017). National Water Initiative. Australian Government, Canberra. Viewed 25 June 2018, http://www.agriculture.gov.au/water/policy/nwi. Department of Agriculture, Water and the Environment, (2019a) Directory of important wetlands in Australia – Information sheet. Limmen Bight (Port Roper) Tidal Wetlands System – NT007. Australian Government, Canberra. Viewed 30 March 2022, https://www.environment.gov.au/cgi-bin/wetlands/report.pl. Department of Agriculture, Water and the Environment, (2019b) Directory of important wetlands in Australia – Information sheet. Mataranka Thermal Pools – NT003. Australian Government, Canberra. Viewed 30 March 2022, https://www.environment.gov.au/cgi- bin/wetlands/report.pl. Department of Agriculture, Water and the Environment (2019c) Species of national environmental significance database. Australian Government, Canberra. Viewed 08 Aug 2019, https://www.awe.gov.au/environment/environmental-information-data/databases- applications/snes. Department of Agriculture, Water and the Environment (2020a) CAPAD2020 Marine. Creative Commons Attribution 3.0 Australian License. Australian Government, Canberra. Viewed 04 September 2021, http://www.environment.gov.au/fed/catalog/search/resource/downloadData.page?uuid=% 7BAF4EE98E-7F09-4172-B95E-067AB8FA10FC%7D. Department of Agriculture, Water and the Environment (2020b) CAPAD2020 Terrestrial. Creative Commons Attribution 3.0 Australian License. Australian Government, Canberra. Viewed 04 September 2021, http://www.environment.gov.au/fed/catalog/search/resource/downloadData.page?uuid=% 7B4448CACD-9DA8-43D1-A48F-48149FD5FCFD%7D. Department of Agriculture, Water and the Environment (2021a) EPBC Act protected matters report: Roper catchment. Australian Government, Canberra. Department of Agriculture, Water and the Environment (2021b) Directory of important wetlands in Australia. Australian Government, Canberra. Available online: https://www.environment.gov.au/water/wetlands/australian-wetlands-database/directory- important-wetlands. Department of the Environment and Energy (2010) A directory of important wetlands in Australia (DIWA) spatial database. [spatial] CSIRO, Canberra. Viewed 08 December 2018, https://www.environment.gov.au/water/wetlands/australian-wetlands-database. Department of Environment and Science Queensland (2013) Coastal and subcoastal floodplain tree swamp – Melaleuca spp. and Eucalyptus spp., WetlandInfo website. Viewed 01 August 2022, https://wetlandinfo.des.qld.gov.au/wetlands/ecology/aquatic-ecosystems- natural/palustrine/floodplain-tree-swamp/. Department of Environment, Parks and Water Security (2018) Water licensing - groundwater and surface water extraction points. Viewed 14/9/2022, https://ftp- dlrm.nt.gov.au/main.html?download&weblink=f75f2af704805fb44b9484017fc19bfd&realfilename=WaterManagement_GW_Licences.zip. Department of Environment, Parks and Water Security (2019a) Fauna atlas N.T. Viewed 01 July 2022. http://www.ntlis.nt.gov.au/metadata/export_data?type=html&metadata_id=2DBCB771208B06B6E040CD9B0F274EFE Department of Environment, Parks and Water Security (2019b) Water control districts in the Northern Territory. Viewed 10 October 2022, https://ftp- dlrm.nt.gov.au/main.html?download&weblink=139913360c21f560ae66df0b5c113205&realfilename=WaterManagement_ControlDistricts.zip. Department of Environment, Parks and Water Security (2000) Melaleuca survey of the Northern Territory. ANZLIC Identifier: 2DBCB7711FE606B6E040CD9B0F274EFE, http://www.ntlis.nt.gov.au/metadata/export_data?metadata_id=2DBCB7711FE606B6E040CD9B0F274EFE&type=html. DeSantis D, James BD, Houchins C, Saur G and Lyubovsky M (2021) Cost of long-distance energy transmission by different carriers. iScience 24(12). https://doi.org/10.1016/j.isci.2021.103495. Dichmont CM, Deng A, Punt AE, Ellis N, Venables WN, Kompas T, Ye Y, Zhou S and Bishop J (2008) Beyond biological performance measures in management strategy evaluation: bringing in economics and the effects of trawling on the benthos. Fisheries Research 94, 238–250. Dostine P and Crook D (2016) Roper River fish movement study. Report on 2015/2016 data. Report number: 19/2016D. Department of Environment and Natural Resources, Palmerston. Douglas MM, Bunn SE and Davies PM (2005) River and wetland food webs in Australia’s wet–dry tropics: general principles and implications for management. Marine and Freshwater Research 56(3), 329–342. Doupe RG, Schaffer J, Knott MJ and Dicky PW (2009) A description of freshwater turtle habitat destruction by feral pigs in tropical north eastern Australia. Herpetological Conservation and Biology 4, 331–339. Duggan M, Bayliss P and Burford MA (2019) Predicting the impacts of freshwater-flow alterations on prawn (Penaeus merguiensis) catches. Fisheries Research 215, 27–37. Duke NC, Kovacs JM, Griffiths AD, Preece L, Hill DJ, Van Oosterzee P, Mackenzie J, Morning HS and Burrows D (2017) Large-scale dieback of mangroves in Australia’s Gulf of Carpentaria: a severe ecosystem response, coincidental with an unusually extreme weather event. Marine and Freshwater Research 68(10), 1816–1829. Earth Sea Heritage Surveys (2013) An assessment of the archaeological potential of Area C and the proposed haul road, Sherwin Iron. Viewed 01 April 2021, https://ntepa.nt.gov.au/__data/assets/pdf_file/0008/290492/sherwin_assessment_report.pdf. Ebner BC, Millington M, Holmes BJ, Wilson D, Sydes T, Bickel TO, Power T, Hammer M, Lach L, Schaffer J, Lymbery A and Morgan DL (2020) Scoping the biosecurity risks and appropriate management relating to the freshwater ornamental aquarium trade across northern Australia. Centre for Tropical Water and Aquatic Ecosystem Research (TropWATER) Publication 20/17. James Cook University, Cairns. Environment Australia (2001) A Directory of Important Wetlands in Australia. Third Edition. Environment Australia, Canberra. Viewed 19 April 2018, http://www.environment.gov.au/system/files/resources/18f0bb21-b67c-4e99-a155- cb5255398568/files/directory.pdf. Faulks JJ (2001) Roper River catchment: an assessment of the physical and ecological condition of the Roper River and its major tributaries. Technical Report No. 36/2001. Natural Resources Division. Department of Lands, Planning and Environment, Katherine, Northern Territory. Finlayson C, Storrs M, Hall R and Bayliss B (1999) Management issues for northern Australian wetlands. In: Finlayson C and Spiers AG (eds) A compendium of information for managing and monitoring wetlands in tropical Australia. Supervising Scientist Report 148. Supervising Scientist, Canberra, 143–167. Finn M and Jackson S (2011) Protecting Indigenous values in water management: a challenge to conventional environmental flow assessments. Ecosystems 14(8), 1232–1248. DOI: 10.1007/s10021-011-9476-0. Gambold N (2015) South East Arnhem Land Indigenous Protected Area – plan of management 2016–2021. Northern Land Council, Darwin. Viewed 31 March 2022, https://www.nlc.org.au/uploads/pdfs/SEAL-IPA-PoM-V3.4.pdf . Garnett ST, Duursma DE, Ehmke G, Guay P-J, Stewart A, Szabo JK, Weston MA, Bennett S, Crowley GM, Drynan D, Dutson G, Fitzherbert K and Franklin DC (2015) Biological, ecological, conservation and legal information for all species and subspecies of Australian bird. Scientific Data 2(1), 150061. DOI: 10.1038/sdata.2015.61. George AK (2004) Eucalypt regeneration on the Lower Murray floodplain, South Australia. PhD Thesis, The University of Adelaide. Georges A, Doody JS, Eisemberg C, Alac EA and Rose M (2008) Carettochelys insculpta Ramsay 1886 – pig-nosed turtle, Fly River turtle. Chelonian Research Monographs, no. 5. Georges A and Thomson SA (2010) Diversity of Australasian freshwater turtles, with an annotated synonymy and keys to species. Zootaxa 2496, 1–37. Geoscience Australia (2017) GEODATA TOPO 250K Series 3 topographic data. Viewed 01 October 2018, http://pid.geoscience.gov.au/dataset/ga/64058. Gleeson J and Richards M (1985) Mataranka and the Daly, two studies in history of settlement in Northern Territory. Northern Research Unit Monograph. Australian National University, Darwin. Godden L, Jackson S and O’Bryan K (2020) Indigenous water rights and water law reforms in Australia. Environmental and Planning Law Journal 37(6), 655–678. Gopurenko D and Hughes JM (2002) Regional patterns of genetic structure among Australian populations of the mud crab, Scylla serrata (Crustacea: Decapoda): evidence from mitochondrial DNA. Marine and Freshwater Research 53(5), 849–857. Gopurenko D, Hughes JM and Bellchambers L (2003) Colonisation of the south-west Australian coastline by mud crabs: evidence for a recent range expansion or human-induced translocation? Marine and Freshwater Research 54(7), 833–840. GPA (2021) Pipelines vs powerlines – a technoeconomic analysis in the Australian context. Report commissioned by the Australian Pipelines and Gas Association. GPA Document No: 210739- REP-001. GPA Engineering, Sydney. Green D, Billy J and Tapim A (2010) Indigenous Australians’ knowledge of weather and climate. Climatic Change 100(2), 337–354. Grey DL, Dall W and Baker A (1983) A guide to the Australian penaeid prawns. Department of Primary Production of the Northern Territory, Darwin. Grill G, Lehner B, Thieme M, Geenen B, Tickner D, Antonelli F, Babu S, Borrelli P, Cheng L, Crochetiere H, Macedo HE, Filgueiras R, Goichot M, Higgins J, Hogan Z, Lip B, McClain ME, Meng J, Mulligan M, Nilsson C, Olden JD, Opperman JJ, Petry P, Liermann CR, Sáenz L, Salinas-Rodríguez S, Schelle P, Schmitt RJP, Snider J, Tan F, Tockner K, Valdujo PH, Soesbergen Av and Zarfl C (2019) Mapping the world’s free-flowing rivers. Nature 569(2019), 215–221. Gunn J (1908) We of the never-never, 15th edition (2003). Macmillan, New York. Hermoso V, Ward DP and Kennard MJ (2013) Prioritizing refugia for freshwater biodiversity conservation in highly seasonal ecosystems. Diversity and Distributions 19(8), 1031–1042. DOI: 10.1111/ddi.12082. Higgins A, McFallan S, Laredo L, Prestwidge D and Stone P (2015) TRANSIT– a model for simulating infrastructure and policy interventions in agriculture logistics: application to the northern Australia beef industry. Computers and Electronics and Agriculture, 114, 32–42. Hill BJ (1975) Abundance, breeding and growth of the crab Scylla serrata in two South African estuaries. Marine Biology 32(2), 119–126. Hill B (1994) Offshore spawning by the portunid crab Scylla serrata (Crustacea: Decapoda). Marine Biology 120(3), 379–384. Hogan A and Nicholson J (1987) Sperm motility of sooty grunter, Hephaestus fuliginosus (Macleay), and jungle perch, Kuhlia rupestris, in different salinities. Marine and Freshwater Research 38(4), 523–528. https://doi.org/10.1071/MF9870523. Integration and Application Network (2022) Integration and Application Network. Attribution- ShareAlike 4.0 International (CC BY-SA 4.0). Bos taurus, Penaeus monodon adult: Jane Hawkey; Corymbia spp, Pritchardia minor, Malaleuca rhaphiophylla, Generic tree rainforest 3, Avicennia germinans, Macropus fuliginosus, Pristis 175ectinate, Chelodina oblonga, Platalea spp, Pluvialis fulva: Tracey Saxby; Pennisetum ciliare, Eucalyptus camaldulensis: Kim Kraeer and Lucy Van Essen-Fishman; Halophila ovata, Halophila beccari: Catherine Collier; Dugong dugong 1, Prawn trawler: Diana Kleine; Lates calcarifer: Dieter Tracey. https://ian.umces.edu/media-library/symbols/#overview. Jackson D, Rolfe JW, English BH, Holmes B, Matthews R, Dixon RM, Smith P and MacDonald N (2012, reprinted 2015) Phosphorus management of beef cattle in northern Australia. Meat and Livestock Australia, Sydney. Jackson S, Finn M, Woodward E and Featherston P (2011) Indigenous socio-economic values and river flows. CSIRO Ecosystem Sciences, Darwin. Jackson S, Finn M and Featherston P (2012) Aquatic resource use by Indigenous Australians in two tropical river catchments: the Fitzroy River and Daly River. Human Ecology 40(6), 893–908. DOI: 10.1007/s10745-012-9518-z. James C, VanDerWal J, Capon S, Hodgson L, Waltham N, Ward D, Anderson B and Pearson R (2013) Identifying climate refuges for freshwater biodiversity across Australia. National Climate Change Adaptation Research Facility, Gold Coast. Jardine TD, Bond NR, Burford MA, Kennard MJ, Ward DP, Bayliss P, Davies PM, Douglas MM, Hamilton SK, Melack JM, Naiman RJ, Pettit NE, Pusey BJ, Warfe DM and Bunn SE (2015) Does flood rhythm drive ecosystem responses in tropical riverscapes? Ecology 96(3), 684–692. Junk W, Bayley P and Sparks R (1989) The flood pulse concept in river-floodplain systems. In: Dodge D (ed.), Proceedings of the International Large River Symposium. Canadian special publication of Fisheries and Aquatic Sciences 106, 110–127. Kennard MJ (2010) Identifying high conservation value aquatic ecosystems in northern Australia. Interim report for the Department of Environment, Water, Heritage and the Arts and the National Water Commission. Charles Darwin University, Darwin. Kennett RM and Russell-Smith J (1993) Seed dispersal by freshwater turtles in northern Australia. In: Lunney D and Ayers D (eds) Herpetology in Australia: a diverse discipline. Royal Zoological Society of New South Wales, Mosman. Kenyon R, Deng R, Plaganyi-Lloyd E and Crous LB (2022) NPF logbook data provided to NAWRA2 Ecology Component, 2022. V1. CSIRO, Data Collection. Csiro:56042. https://data.csiro.au/collection/csiro:56042. King AJ, Townsend SA, Douglas MM and Kennard MJ (2015) Implications of water extraction on the low-flow hydrology and ecology of tropical savannah rivers: an appraisal for northern Australia. Freshwater Science 34(2), 741–758. DOI: 10.1086/681302. Kingsford RT, Porter J and Halse SA (2012) National waterbird assessment. Waterlines Report. National Water Commission, Canberra. Laird A (2021) Northern Prawn Fishery data summary 2020. NPF Industry Pty Ltd, Australia. Available online: https://www.afma.gov.au/sites/default/files/2023- 02/npf_data_summary_2020.pdf https://www.afma.gov.au/sites/default/files/npf_data_summary_2020.pdf. Larsen A, May JH, Moss P and Hacker J (2016) Could alluvial knickpoint retreat rather than fire drive the loss of alluvial wet monsoon forest, tropical northern Australia? Earth Surface Processes and Landforms 41(11), 1583–1594. Lear KO, Gleiss AC, Whitty JM, Fazeldean T, Albert J, Green N, Ebner BC, Thorburn DC, Beatty SJ and Morgan DL (2019) Recruitment of a critically endangered sawfish into a riverine nursery depends on natural flow regimes. Scientific Reports 9(1), 1–11. Legler JM (1982) The ecology of freshwater turtles in the Alligator Rivers Region. Unpublished report to the Office of the Supervising Scientist, Department of Employment and Industrial Relations, Canberra. Loneragan N, Bunn SE and Kellaway D (1997) Are mangroves and seagrasses sources of organic carbon for penaeid prawns in a tropical Australian estuary? A multiple stable-isotope study. Marine Biology 130(2), 289–300. Lucas C, Kirkwood G and Somers I (1979) An assessment of the stocks of the banana prawn Penaeus merguiensis in the Gulf of Carpentaria. Marine and Freshwater Research 30(5), 639–652. Lyons P, Barber M, Fisher K and Braedon P (2023) Indigenous water values, rights, interests and development goals in the Roper catchment. A technical report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Macintosh, A. Waschka, M. Jones, J. Wood, A. (2018) Legal, regulatory and policy environment for the development of water resources in northern Australia. A technical report from the CSIRO Northern Australia Water Resource Assessment to the Government of Australia. CSIRO, Australia Marchant S and Higgins P (1990) Handbook of Australian, New Zealand and Antarctic birds. Volume 1, Ratites to ducks. Oxford University Press, Melbourne. McGinness H, Brandis K, Robinson F, Piper M, O’Brien L, Langston A, Hodgson J, Wenger L, Martin J and Bellio M (2019) Murray–Darling Basin environmental water knowledge and research project: waterbirds theme research report. Report prepared for the Department of the Environment and Energy, Commonwealth Environmental Water Office by CSIRO and La Trobe University, Centre for Freshwater Ecosystems, CFE Publication 225. CSIRO Canberra, Australia. McGrath A (1987) Born in the cattle: Aborigines in cattle country. Allen and Unwin, Sydney. McIntyre-Tamwoy S, Bird M and Akinson F (2013) Aboriginal cultural heritage in the Flinders and Gilbert rivers: a desktop review – report to the CSIRO by Archaeological and Heritage Management Solutions. Archaeological and Heritage Management Solutions, Sydney. McJannet D, Wallace J, Henderson A and McMahon J (2009) High and low flow regime changes at environmental assets across northern Australia under future climate and development scenarios. A report to the Australian Government from the CSIRO Northern Australia Sustainable Yields Project. CSIRO, Australia. McJannet D, Marvanek S, Kinsey-Henderson A, Petheram C and Wallace J (2014) Persistence of in- stream waterholes in ephemeral rivers of tropical northern Australia and potential impacts of climate change. Marine and Freshwater Research 65(12), 1131–1144. DOI: 10.1071/MF14035. Merlan F (1978) ‘Making people quiet’ in the pastoral north: reminiscences of Elsey Station. Aboriginal History 2(1), 70–106. Merlan F (1981) Land, language and social identity in Aboriginal Australia. Mankind 13(2), 133– 148. Merlan F (1982) A Mangarrayi representational system: environmental and cultural symbolization in northern Australia. American Anthropologist 9, 145–166. Merlan F (1986) Mataranka land claim – senior anthropologist’s report on behalf of the claimants. Darwin. Meynecke J-O and Lee SY (2011) Climate-coastal fisheries relationships and their spatial variation in Queensland, Australia. Fisheries Research 110(2), 365–376. Meynecke J, Lee S, Grubert M, Brown I, Montgomery S, Gribble N, Johnston D and Gillson J (2010) Evaluating the environmental drivers of mud crab (Scylla serrata) catches in Australia. Final report 2002/012. The Fisheries Research and Development Corporation and Griffith University. Mitchell PJ, O’Grady AP, Pinkard EA, Brodribb TJ, Arndt SK, Blackman CJ, Duursma RA, Fensham RJ, Hilbert DW and Nitschke CR (2016) An ecoclimatic framework for evaluating the resilience of vegetation to water deficit. Global Change Biology 22(5), 1677–1689. Mitsch WJ, Bernal B and Hernandez ME (2015) Ecosystem services of wetlands. International Journal of Biodiversity Science, Ecosystem Services & Management, 11(1), 1–4. DOI: 10.1080/21513732.2015.1006250. Morphy H and Morphy F (1981) Yutpundji-Djindiwirritj (Roper Bar) land claim book. Northern Land Council, Darwin. Murphy BP and Bowman DM (2012) What controls the distribution of tropical forest and savanna? Ecology Letters 15(7), 748–758. NAILSMA (2008) Garma International Indigenous Water Declaration. In: Garma Conference. Gulkula, Northern Territory. North Australian Indigenous Land and Sea Management Alliance Ltd, Darwin. NAILSMA (2009) Mary River Statement. North Australian Indigenous Land and Sea Management Alliance Ltd, Darwin. NAILSMA (2012) Towards resilient communities through reliable prosperity. North Australian Indigenous experts forum on sustainable economic development – first forum report, Mary River Park, Northern Territory, 19–21 June 2012. North Australian Indigenous Land and Sea Management Alliance Ltd, Darwin. NAILSMA (2013) An Indigenous prospectus for participating in the sustainable development of north Australia. North Australian Indigenous experts forum on sustainable economic development – second forum report, Kakadu National Park, Northern Territory, 30 April – 2 May 2013. North Australian Indigenous Land and Sea Management Alliance Ltd, Darwin. Naughton JM, O’Dea K and Sinclair AJ (1986) Animal foods in traditional Australian aboriginal diets: polyunsaturated and low in fat. Lipids 21(11), 684–690. Nikolakis W and Grafton RQ (2022) Law versus justice: the Strategic Aboriginal Water Reserve in the Northern Territory, Australia. International Journal of Water Resources Development 38(1), 11–29. DOI: 10.1080/07900627.2021.1882406. Nielsen DL, Cook RA, Ning N, Gawne B and Petrie R (2015) Carbon and nutrient subsidies to a lowland river following floodplain inundation. Marine and Freshwater Research 67(9), 1302– 1312. Northern Territory Government (2017) Strategic Aboriginal Water Reserve policy framework. Department of Environment and Natural Resources, Darwin. Northern Territory Government (2018) Status of key Northern Territory fish stocks report 2016. Fishery Report No 119. Department of Primary Industry and Resources, Darwin. Available online: https://dpir.nt.gov.au/__data/assets/pdf_file/0007/546640/FR119.pdf. Northern Territory Government (2020) Important wetlands. Northern Territory Government, Darwin. Available online: https://nt.gov.au/environment/soil-land-vegetation/important- wetlands. NT PHN (Northern Territory Primary Health Network) (2020) Katherine region data report: overview of selected demographic and health data for the Katherine region of the Northern Territory. Australian Government, Department of Health and Aged Care, Darwin. Oberprieler S, Rees G, Nielsen D, Shackleton M, Watson G, Chandler L and Davis J (2021) Connectivity, not short-range endemism, characterises the groundwater biota of a northern Australian karst system. Science of the Total Environment 796, 148955. O’Donnell E, Jackson S, Langton M and Godden L (2022) Racialized water governance: the ‘hydrological frontier’ in the Northern Territory, Australia. Australasian Journal of Water Resources 26(1), 59–71. DOI: 10.1080/13241583.2022.2049053. Palmer C and Smit N (2019) Limmen Bight Marine Park: marine and coastal biodiversity values. Department of Environment and Natural Resources, Darwin. Peterson N (2013) On the persistence of sharing: personhood, asymmetrical reciprocity, and demand sharing in the Indigenous Australian domestic moral economy. The Australian Journal of Anthropology 24(2), 166–176. DOI: 10.1111/taja.12036. Pettit C (undated) Sturt Plateau District. Land condition guide. Understanding the productivity of grazing lands. Northern Territory Government. Pettit NE, Bayliss P, Davies PM, Hamilton SK, Warfe DM, Bunn SE and Douglas MM (2011) Seasonal contrasts in carbon resources and ecological processes on a tropical floodplain. Freshwater Biology 56(6), 1047–1064. DOI: 10.1111/j.1365-2427.2010.02544.x. Pettit NE, Jardine TD, Hamilton SK, Sinnamon V, Valdez D, Davies PM, Douglas MM and Bunn SE (2012) Seasonal changes in water quality and macrophytes and the impact of cattle on tropical floodplain waterholes. Marine and Freshwater Research 63(9), 788–800. DOI: 10.1071/MF12114. Pettit N, Dowe J and Dixon I (2016) Riparian vegetation of tropical Northern Australia. In: Capon S, James C and M R (eds) Vegetation of Australian riverine landscapes: biology, ecology and management. CSIRO, Australia, 221–237. Pettit NE, Naiman RJ, Warfe DM, Jardine TD, Douglas MM, Bunn SE and Davies PM (2017) Productivity and connectivity in tropical riverscapes of Northern Australia: ecological insights for management. Ecosystems 20(3), 492–514. Pidgeon R and Humphrey C (1999) Some important ecological features of wetlands in the wet-dry tropics of northern Australia. In: Finlayson C and Spiers AG (eds) A compendium of information for managing and monitoring wetlands in tropical Australia. Supervising Scientist Report 148. Supervising Scientist, Canberra, 40–62. Poff NL, Allan JD, Bain MB, Karr JR, Prestegaard KL, Richter BD, Sparks RE and Stromberg JC (1997) The natural flow regime: a paradigm for river conservation and restoration. BioScience 47(11), 769–784. Poiner I, Staples D and Kenyon R (1987) Seagrass communities of the Gulf of Carpentaria, Australia. Marine and Freshwater Research 38(1), 121–131. Pollino CA, Barber E, Buckworth R, Deng A, Ebner B, Kenyon R, Liedloff A, Merrin LE, Moeseneder C, Nielsen DL, O’Sullivan J, Ponce Reyes R, Robson BJ, Stratford DS, Stewart- Koster B and Turschwell M (2018b) Synthesis of knowledge to support the assessment of impacts of water resource development to ecological assets in northern Australia: asset analysis. A technical report to the Australian Government from the CSIRO. Northern Australia Water Resource Assessment, part of the National Water Infrastructure Development Fund: Water Resource Assessments. CSIRO, Australia. Pusey BJ, Kennard MJ and Arthington AH (2004) Freshwater fishes of north-eastern Australia. CSIRO Publishing, Australia. Ramsar Convention Secretariat (2004) The Ramsar Convention manual: a guide to the Convention on Wetlands (Ramsar, Iran, 1971). Third edition. Ramsar Convention Secretariat, Gland, Switzerland. Robins JB, Northrop AR, Grubert M and Buckworth RC (2020) Understanding environmental and fisheries factors causing fluctuations in mud crab and blue swimmer crab fisheries in northern Australia to inform harvest strategies. Final report to the Fisheries Research and Development Corporation for Project Number 2017/047. Department of Agriculture and Fisheries, Brisbane. Rogers K and Ralph TJ (2010) Floodplain wetland biota in the Murray-Darling Basin: water and habitat requirements. CSIRO Publishing, Australia. Rose D (2002) Country of the heart – an Indigenous Australian homeland. Aboriginal Studies Press, Canberra. Russell D and Garrett R (1983) Use by juvenile barramundi, Lates calcarifer (Bloch), and other fishes of temporary supralittoral habitats in a tropical estuary in northern Australia. Marine and Freshwater Research 34(5), 805–811. Russell D and Garrett R (1985) Early life history of barramundi, Lates calcarifer (Bloch), in north- eastern Queensland. Marine and Freshwater Research 36(2), 191–201. Russell S, Ens E and Ngukurr Yangbala Rangers (2021) Diving into Indigenous biocultural knowledge of freshwater turtles in South East Arnhem Land. News, 22 December 2021. Atlas of Living Australia. Russell-Smith J (1991) Classification, species richness, and environmental relations of monsoon rain forest in northern Australia. Journal of Vegetation Science 2(2), 259–278. Ryan L, Debenham J, Pascoe B, Gilbert S, Smith R, Brown M, Price D and Newley J (2018) Colonial Frontier Massacres in Central and Eastern Australia 1788-1930. University of Newcastle. Viewed 23 March 2023, . Sandefur J (1985) Aspects of the socio-political history of Ngukurr (Roper River) and its effect on language change. Aboriginal History 9(2), 205–219. Savage J and Hobsbawn P (2015) Australian fisheries and aquaculture statistics 2014. Fisheries Research and Development Corporation project 2014/245. ABARES, Canberra. Schofield KA, Alexander LC, Ridley CE, Vanderhoof MK, Fritz KM, Autrey BC, DeMeester JE, Kepner WG, Lane CR, Leibowitz SG and Pollard AI (2018) Biota connect aquatic habitats throughout freshwater ecosystem mosaics. Journal of the American Water Resources Association 54(2), 372–399. SKM (2009) Ecological assets of northern Australia study. Sinclair Knight Mertz, Canberra. Smott S, Monczak A, Miller ME and Montie EW (2018) Boat noise in an estuarine soundscape – a potential risk on the acoustic communication and reproduction of soniferous fish in the May River, South Carolina. Marine Pollution Bulletin 133, 246–260. https://doi.org/10.1016/j.marpolbul.2018.05.016. Smyth C and Turner J (2019) Natural values and resource use in the Limmen Bight Marine Region. Australian Marine Conservation Society, West End. Stanford C, Iverson J, Rhodin A, van Dijk P, Mittermeier RK, G, Berry K, Bertolero A, Bjorndal K, Blanck T, Buhlmann K, Burke R, Congdon J, Diagne T, Edwards T, Eisemberg C, Ennen J, Forero-Medina G, Frankel M, Fritz U, Gallego-García N, Georges A, Gibbons J, Gong S, Goode E, Shi H, Hoang H, Hofmeyr M, Horne B, Hudson R, Juvik J, Kiester R, Koval P, Le M, Lindeman P, Lovich J, Luiselli L, McCormack T, Meyer G, Páez V, Platt K, Platt S, Pritchard P, Quinn H, Roosenburg W, Seminoff J, Shaffer H, Spencer R, Van Dyke J, Vogt R and Walde A (2020) Turtles and tortoises are in trouble. Current Biology 30(12), R721–R735. Staples D (1979) Seasonal migration patterns of postlarval and juvenile banana prawns, Penaeus merguiensis de Man, in the major rivers of the Gulf of Carpentaria, Australia. Marine and Freshwater Research 30(2), 143–157. Staples D (1980) Ecology of juvenile and adolescent banana prawns, Penaeus merguiensis, in a mangrove estuary and adjacent off-shore area of the Gulf of Carpentaria. I. Immigration and settlement of postlarvae. Marine and Freshwater Research 31(5), 635–652. Staples D, Vance D and Heales D (1985) Habitat requirements of juvenile penaeid prawns and their relationship to offshore fisheries. National Prawn Seminar 2, Cleveland, Queensland. Steven AH, Dylewski M and Curtotti R (2021) Australian fisheries and aquaculture statistics 2020. Fisheries Research and Development Corporation, ABARES, Canberra. https://doi.org/10.25814/0wzy-re76. Stokes C, Jarvis D, Webster A, Watson I, Jalilov S, Oliver Y, Peake A, Peachey A, Yeates S, Bruce C, Philip S, Prestwidge D, Liedloff A, Poulton P, Price B and McFallan S (2023) Financial and socio-economic viability of irrigated agricultural development in the Roper catchment. A technical report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Stork NE, Goosem S and Turton SM (2008) Australian rainforests in a global context. In: Stork NE and Turton SM (eds) Living in a dynamic tropical forest landscape. Wiley-Blackwell, 4–20. Strang V (1997) Uncommon ground: cultural landscapes and environmental values. Berg, Oxford, New York. Stratford D, Kenyon R, Pritchard J, Merrin L, Linke S, Ponce Reyes R, Blamey L, Buckworth R, Castellazzi P, Costin B, Deng R, Gannon R, Gilbey S, King D, Kopf K, Kopf S, McGinness H, McInerney P, Perna C, Plaganyi E and Waltham N (2022) Ecological assets of northern Australia to inform water resource assessments. A technical report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Tanimoto M, Robins JB, O’Neill MF, Halliday IA and Campbell AB (2012) Quantifying the effects of climate change and water abstraction on a population of barramundi (Lates calcarifer), a diadromous estuarine finfish. Marine and Freshwater Research 63(8), 715–726. Taylor AR, Crosbie RS, Turnadge C, Lamontagne S, Deslandes A, Davies PJ, Barry K, Suckow A, Knapton A, Marshall S, Hodgson G, Tickell S, Duvert C, Hutley L and Dooley K (2023) Hydrogeological assessment of the Cambrian Limestone Aquifer and the Dook Creek Aquifer in the Roper catchment, Northern Territory. A technical report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Taylor AR, Crosbie RS, Turnadge C, Lamontagne S, Deslandes A, Davies PJ, Barry K, Suckow A, Knapton A, Marshall S, Hodgson G, Tickell S, Duvert C, Hutley L and Dooley K (2023) Hydrogeological assessment of the Cambrian Limestone Aquifer and the Dook Creek Aquifer in the Roper catchment, Northern Territory. A technical report from the CSIRO Roper River Water Resource Assessment for the National Water Grid. CSIRO, Australia. Thomson SA (2000) The identification of the holotype of Chelodina oblonga (Testudines: Chelidae) with a discussion of taxonomic implications. Chelonian Conservation Biology 3(4), 745–749. Thomson SA, Kennet R and Georges A (2000) A new species of long-necked turtle (Testudines: Chelidae) from the Arnhem Land Plateau, Northern Territory, Australia. Chelonian Conservation and Biology 3, 675–685. Tomkovish PS and Weston MA (2008) Breeding ecology. In: Geering A, Agnew L and Harding S (eds), Shorebirds of Australia. CSIRO Publishing, Australia. United Nations (2023) Indigenous Peoples’ Declaration for the 2023 United Nations Water Conference. Viewed 15 May 2023, https://www.un.org/development/desa/indigenouspeoples/news/2023/01/2023-unwater/ van Dam R, Bartolo R and Bayliss P (2008) Identification of ecological assets, pressures and threats. Environmental Research Institute of the Supervising Scientist, Darwin, 14–161. van Dijk P, Iverson JB, Rhodin AG, Shaffer HB and Bour RJ, (eds) (2014) Turtles of the world, seventh edition: annotated checklist of taxonomy, synonymy, distribution with maps, and conservation status. Chelonian Research Monographs, no. 5. Vance D, Staples D and Kerr J (1985) Factors affecting year-to-year variation in the catch of banana prawns (Penaeus merguiensis) in the Gulf of Carpentaria, Australia. ICES Journal of Marine Science 42(1), 83–97. Vance D, Haywood M and Staples D (1990) Use of a mangrove estuary as a nursery area by postlarval and juvenile banana prawns, Penaeus merguiensis de Man, in northern Australia. Estuarine, Coastal and Shelf Science 31(5), 689–701. Vance D, Haywood M, Heales D, Kenyon R and Loneragan N (1998) Seasonal and annual variation in abundance of postlarval and juvenile banana prawns Penaeus merguiensis and environmental variation in two estuaries in tropical northeastern Australia: a six year study. Marine Ecology Progress Series 163, 21–36. Vance D, Bishop J, Dichmont C, Hall N, McInnes K and Taylor B (2003) Management of common banana prawn stocks of the Gulf of Carpentaria: separating the effects of fishing from those of the environment. CSIRO Report 1998/0716. CSIRO, Brisbane. Vörösmarty CJ, McIntyre PB, Gessner MO, Dudgeon D, Prusevich A, Green P, Glidden S, Bunn SE, Sullivan CA, Liermann CR and Davies PM (2010) Global threats to human water security and river biodiversity. Nature 467(7315), 555–561. DOI: 10.1038/nature09440. Waltham N, Burrows D, Butler B, Wallace J, Thomas C, James C and Brodie J (2013) Waterhole ecology in the Flinders and Gilbert catchments. A technical report to the Australian Government from the CSIRO Flinders and Gilbert Agricultural Resource Assessment, part of the North Queensland Irrigated Agriculture Strategy. CSIRO Water for a Healthy Country and Sustainable Agriculture flagships, Australia. Ward JV and Stanford J (1995) Ecological connectivity in alluvial river ecosystems and its disruption by flow regulation. Regulated Rivers: Research & Management 11(1), 105–119. Warfe DM, Pettit NE, Davies PM, Pusey BJ, Hamilton SK, Kennard MJ, Townsend SA, Bayliss P, Ward DP, Douglas MM, Burford MA, Finn M, Bunn SE and Halliday IA (2011) The ‘wet-dry’ in the wet-dry tropics drives river ecosystem structure and processes in northern Australia. Freshwater Biology 56(11), 2169–2195. DOI: 10.1111/j.1365-2427.2011.02660.x. Welch D, Robins J and Saunders T (2014) Implications of climate change impacts on fisheries resources of northern Australia. Part 2: species profiles. Fisheries Research Development Corporation, Final report project 2010/565. James Cook University, Townsville. Wen L, Ling J, Saintilan N and Rogers K (2009) An investigation of the hydrological requirements of river red gum (Eucalyptus camaldulensis) forest, using classification and regression tree modelling. Ecohydrology 2(2), 143–155. Williams N (1986) The Yolngu and their land: a system of land tenure and the fight for its recognition. Australian Institute of Aboriginal Studies, Canberra. Woinarski JC (2004) The forest fauna of the Northern Territory: knowledge, conservation and management. In: Lunney D (ed.) Conservation of Australia’s forest fauna. Royal Zoological Society of New South Wales, Mosman. Woodhams J, Larcombe J and George D (2011) Northern Prawn Fishery. In: Woodhams J, Stobutzki I, Vieira S, Curtotti R and Begg GA (eds) Fishery status reports 2010: status of fish stocks and fisheries managed by the Australian Government. ABARES, Canberra. World Water Council (2003) Indigenous peoples Kyoto water declaration. Viewed 01 October 2022, https://www.worldwatercouncil.org/en/kyoto-2003 http://www.waterculture.org/uploads/IPKyotoWaterDeclarationFINAL.pdf. Xiang H, Zhang Y and Richardson JS (2016) Importance of riparian zone: effects of resource availability at land-water interface. Riparian Ecology and Conservation 2016(3), 1–17. Zaar U (2009) Gulf water study: Roper River region. Report 16/2009D. Northern Territory Government, Darwin. Zhou S, Buckworth RC, Ellis N, Deng RA and Pascoe S (2015) Getting all information out of logbooks: estimating banana prawn fishable biomass, catchability, and fishing power increase, with a focus on natural mortality. ICES Journal of Marine Science 72(1), 54–61. Zoellner D (2017) Vocational education & training. The Northern Territory’s history of public philanthropy. ANU Press, Canberra.